Over the last 50 years, the theory of rational choice has

advertisement
ARBITRAGE, INCOMPLETE MODELS, AND INTERACTIVE
RATIONALITY
Robert F. Nau
Fuqua School of Business
Duke University
Durham, NC 27708-0120 USA
robert.nau@duke.edu
www.duke.edu/~rnau
(919) 660-7763
Version 3.5
October 13, 1999
ABSTRACT: Rational choice theory rests on the assumption that decision makers have
complete mental models of the events, consequences and acts that constitute their environment.
They are considered to be individually rational if they hold preferences among acts that satisfy
axioms such as ordering and independence, and they are collectively rational if they satisfy
additional postulates of inter-agent consistency such as common knowledge and common prior
beliefs. It follows that rational decision makers are expected-utility maximizers who dwell in
conditions of equilibrium. But real decision makers are only boundedly rational, they must cope
with disequilibrium and environmental change, and their decision models are incomplete. As
such, they are often unable or unwilling to behave in accordance with the rationality axioms, they
find the theory hard to apply to most personal and organizational decisions, and they regard the
theory’s explanations of many economic and social phenomena to be unsatisfying. Models and
experimental studies of bounded rationality, meanwhile, often focus on the behavior of unaided
decision makers who employ strategies such as satisficing or adaptive learning that can be
implemented with finite attention, memory, and computational ability.
This essay proposes a new foundation for choice theory that does not rely on consequences, acts,
and preferences as primitive concepts. Rather, agents articulate their beliefs and values through
the acceptance of small gambles or trades in a stylized market. These primitive measurements
are intersubjective in nature, eliminating the need for separate common knowledge assumptions,
and they partly endogenize the writing of the “rules of the game.” In place of the assorted
preference axioms and inter-agent consistency conditions of the standard theory, only a single
axiom of substantive rationality is needed, namely the principle of no arbitrage. No-arbitrage is
shown to be the primal characterization of rationality with respect to which solution concepts
such as expected-utility maximization and strategic and competitive equilibria are merely dual
characterizations. The traditional distinctions among individual, strategic, and competitive
rationality are thereby dissolved.
Arbitrage choice theory (ACT) does not require the decision models of individuals to be
complete, so it is compatible with the notion that individual rationality is bounded. It is also
inherently a theory of group rationality rather than individual rationality, so it can be applied at
any level of activity from personal decisions to games of strategy to competive markets. This
group-centered view of rationality admits the possibility that individuals do more than merely
satisfice when making decisions in complex environments for which they lack complete models.
Rather, they use “other people’s brains” by seeking advice from colleagues and experts, forming
teams or management hierarchies, consulting the relevant literature, relying on market prices,
invoking social norms, and so on. The most important result of such interactive decision
processes usually is not the identification of an existing alternative that optimizes the latent
beliefs and values of the putative decision maker, but rather the synthesis of a more complete
model of the problem than she (or perhaps anyone) initially possesses, the construction of sharper
beliefs and values, and the discovery or creation of new (and perhaps dominant) alternatives. On
this view, traditional models of rational choice that attempt to explain the behavior of
households, firms, markets, and polities purely in terms of the satisfaction of individual
preferences are perhaps overlooking the most important purpose of socioeconomic interactions,
namely that they harness many people’s brains to the solution of complex problems.
ii
Contents
1.1 Introduction
2.1 Outline of standard rational choice theory
1. Environment
2. Behavior
3. Rationality
2.2 Trouble in paradise
1. Consequences and acts
2. Preferences
3. The axioms
4. Equilibrium
5. Impossibility
6. Simple questions, equivocal answers
7. The fossil record
2.3 Alternative paradigms
1. Bounded rationality, behavioral economics, and organization theory
2. Behavioral decision theory and experimental economics
3. Austrian and subjectivist economics
4. Evolutionary & complexity theory
3.1 Outline of arbitrage choice theory
1. Environment
2. Behavior
3. Rationality
3.2 Fundamental theorem and examples
1. Pure exchange
2. Elicitation and aggregation of belief
3. Decisions under uncertainty
4. Games of strategy
5. Learning from experience
6. Allais’ paradox
3.3 On the realism and generality of the modeling assumptions
3.4 Summary
4.1 Incomplete models and other people’s brains
4.2 The limits of theory
4.3 Implications for Modeling
iii
Arbitrage, Incomplete Models, and Interactive Rationality
Robert F. Nau
1.1 Introduction
Over the last 50 years, the theory of rational choice has emerged as the dominant paradigm of
quantitative research in the social and economic sciences. The idea that individuals make
choices by rationally weighing values and uncertainties (or that they ought to, or at least act as if
they do) is central to Bayesian methods of statistical inference and decision analysis; the theory
of games of strategy; theories of competitive markets, industrial organization, and asset pricing;
the theory of social choice; and a host of rational actor models in political science, sociology,
law, philosophy, and management science.
Rational choice theory is founded on the principles of methodological individualism and
purposive action. Methodological individualism means that social and economic phenomena are
explained in terms of “a particular configuration of individuals, their dispositions, situations,
beliefs, and physical resources and environment” rather than in terms of holistic or emergent
properties of groups. (Watkins 1957; c.f. Homans 1967, Brodbeck 1968, Ordeshook 1986,
Coleman 1990, Arrow 1994) Purposive action means that those individuals have clear and
consistent objectives and they employ reason to find the best means of achieving those
objectives. They intend their behavior to cause effects that they desire, and their behavior does
cause those effects for the reasons they intend. (Ordeshook 1986, Elster 1986) In most versions
of the theory, the objectives of the individuals are expressed in terms of preferences: they choose
what they most prefer from among the alternatives available, and the (only) role of social and
economic institutions is to enable them to satisfy their preferences through exchange or strategic
contests with other individuals.
This essay sketches the outline of arbitrage choice theory (ACT), a new synthesis of rational
choice that weakens the emphasis on methodological individualism and purposive action and
departs from the traditional use of preference as a behavioral primitive, building on earlier work
by Nau and McCardle (1990, 1991) and Nau (1992abc, 1995). The main axiom of rationality in
this framework is the requirement of no arbitrage, and it leads to a strong unification of the
theories of personal decisions, games of strategy, and competitive markets. It also yields a very
different perspective on the purpose of social and economic interactions between agents,
suggesting that group behavior is to some extent emergent and suprarational, not merely an
aggregation or collision of latent individual interests. Thus, we will take issue with the following
statement of Elster (1986): “A family may, after some discussion, decide on a way of spending
its income, but the decision is not based on ‘its’ goals and ‘its’ beliefs, since there are no such
things.” We will argue that there are indeed “such things” and that they may be better defined for
the group than for the constituent individuals.
The organization is as follows. Section 2 presents an outline and brief critique of the standard
rational choice theory. Section 3 presents a contrasting outline of arbitrage choice theory choice,
1
illustrated by a few simple examples. Section 4 focuses on the phenomenon of model
incompleteness and its implications for interactions between agents and the emergence of grouplevel rationality.
2.1 Outline of the standard rational choice theory
Standard rational choice theory begins with the assumption that the infinitely detailed “grand
world” in which real choices are made can be adequately approximated by a “small world” model
with a manageable number of numerical parameters. The formal theory is then expressed in
terms of assumptions about the structure of the small-world environment, the modes of behavior
that take place in that environment, and the conditions that behavior must satisfy in order to
qualify as rational.
Elements of standard rational choice theory
1. Environment:
Agents
Events (states of nature and alternatives for agents)
Consequences
Mappings of events to consequences for all agents
Acts (hypothetical mappings of events not under an agent’s control to
consequences)
2. Behavior:
Physical behavior: choices among alternatives
Mental behavior: preferences among acts
3. Rationality:
(a) Individual rationality
Axioms of preference (completeness, transitivity, independence, etc.)
Axiom of rational choice (choices agree with preferences)
(b) Strategic rationality
Common knowledge of individual rationality
Common knowledge of utilities
Common prior probabilities
Probabilistic independence
(c) Competitive rationality
Price-taking
Market clearing
(d) Social rationality
Pareto efficiency
(e) Rational asset pricing
No arbitrage
(f) Rational learning
Bayes’ theorem
(g) Rational expectations
Self-fulfilling beliefs
2
1. Environment: The environment is inhabited by one or more human agents (also called
actors or players), and in the environment events may happen, under the control of the agents
and/or nature. An event that is under an agent’s control is an alternative, and an event that is
under nature’s control is a state of nature. States of nature just happen, while alternatives are
chosen. A realization of events is called an outcome. For each agent there is a set of material or
immaterial consequences (wealth, health, pleasure, pain, etc.) that she may enjoy or suffer, and
there is a known mapping from outcomes to consequences.1 For example, an agent may face a
choice between the alternatives “walk to work with an umbrella” or “walk to work without an
umbrella,” the states of nature might be “rain” or “no rain,” and the consequences might be “get
slightly wet” or “get drenched” or “stay dry,” with or without the “hassle” of carrying an
umbrella, as summarized in the following contingency table:
Table 1
State of nature:
Alternative:
Walk to work with umbrella (a1)
Walk to work without umbrella (a2)
Rain (s1)
Get slightly wet, hassle (c1)
Get drenched, no hassle (c3)
No rain (s2)
Stay dry, hassle (c2)
Stay dry, no hassle (c4)
From the perspective of an observer of the situation, every cell in the table corresponds to an
outcome: nature will or won’t rain and the agent will or won’t carry her umbrella. Every
outcome, in turn, yields a known consequence. From the perspective of the agent, every row in
the table corresponds to an alternative, and her problem is to choose among the alternatives.
Here the alternatives have been labeled as a1, a2; the states of nature have been labeled as s1, s2;
and the consequences have been labeled as c1 through c4.
Every alternative for an agent is a feasible mapping of events not under that agent’s control to
consequences. An act for an agent is an arbitrary mapping of events not under that agent’s
control to consequences. Thus, an act is a (usually) hypothetical alternative, while an alternative
is a feasible act. For example, the act “take a taxicab to work at a cost of $5” might yield the
consequence “ride in comfort minus $5” (henceforth labeled c5) whether it rains or not, and an
agent can contemplate such an act regardless of whether it is feasible. (The cab drivers may be
on strike today, but the agent still can imagine the ride.) The set of acts is typically much richer
than the set of alternatives for each agent. For example, the set of acts for our protagonist with
the umbrella might be represented by the following table:
The terminology used here is slightly non-standard. Normally the term “event” refers to a set of states of nature,
and the term “outcome” is sometimes synonymous with consequence. Here the term “event” refers generically to a
set of states of nature and/or choices by human agents, and an outcome is an atomic event—e.g., a single cell in a
contingency table. For an agent, every outcome is mapped to a known consequence, but different outcomes might, in
principle, lead to the “same” consequence for that agent. The generic use of the term “event” admits the possibility
that the objects of one agent’s uncertainty may be acts of other agents as well as acts of nature. Meanwhile, the term
“alternative” is used here to refer to what is sometimes called a “concrete act,” i.e., a feasible object of choice for a
human agent.
1
3
Table 2
Act:
a1
a2
a3
...
ai
...
State of nature:
s1 s2
c1 c2
c3 c4
c5 c5
... ...
cj ck
... ...
Here, acts a1 and a2 happen to correspond to alternatives a1 and a2 (walking with or without the
umbrella) while act a3 (taking the cab) might be purely hypothetical. Another act (ai) might be
composed of an arbitrary assignment of consequences (say cj and ck) to weather states—even an
oxymoron such as “stay dry, no hassle” if it rains, “get drenched, hassle” if it doesn’t rain. The
small world is therefore rather “big,” since it contains a great number of infeasible alternatives in
addition to the feasible ones.
2. Behavior: Within the environment, several kinds of behavior occur. First and most
importantly, there is physical behavior, which affects the outcomes of events. Physical behavior
by agents2 consists of choices among feasible alternatives, corresponding to the selection of a
single row out of a table similar to Table 1. However, it does not suffice to model only physical
behavior, because the set of feasible alternatives usually is not rich enough to support a tight
mathematical representation and because it is of interest to predict choices from other kinds of
antecedent behavior. The other kind of behavior most often modeled in rational choice theory is
preference behavior. Preferences are hypothetical choices between hypothetical alternatives
(acts), corresponding to the selection of one row (or perhaps an equivalence class of several
rows) out of a table similar to Table 2. An agent prefers act x to act y if she imagines that she
would choose x rather than y if given the choice between them. (Actually, we are getting ahead
of ourselves: at this point preference is merely an undefined primitive, but it will be linked to
choices in the axioms which follow.) Preference behavior may be interpreted as a kind of
mental behavior that precedes and ultimately causes physical behavior. The agent is assumed to
have preferences with respect to all possible acts, even those that involve counterfactual
assumptions, such as: “If I had a choice between riding in a cab (which is not running today) or
walking to work with an umbrella (which I don’t have), I would take the cab.” The domain of
preferences is therefore very rich, providing the quantitative detail needed to support a finegrained mathematical representation of behavior.
3. Rationality: Assumptions about what it means to behave rationally are typically imposed at
several levels: the level of the agent (individual rationality), the level of the small group
(strategic rationality), and the level of the large group (competitive or social rationality). The
Nature also “behaves” by choosing among the states, but nature’s moves are modeled differently than those of the
human agents. In the alternative theory introduced in section 3, states of nature and alternatives of agents are treated
more symmetrically.
2
4
lowest (agent) level assumptions apply in all cases, while the higher (group) level assumptions
are applied in different combinations depending on the situation.
a. Individual rationality. The principal assumptions of individual rationality are axioms
imposed on mental behavior—that is, on preferences. Preferences are assumed to establish an
ordering of acts from least-preferred to most-preferred, and they are usually taken to satisfy the
axioms most often imposed on ordering relations so as to ensure the existence of a convenient
numerical representation. For example, preferences are usually assumed to be complete, so that
for any acts x and y an agent always knows whether she prefers x to y or y to x or is precisely
indifferent between the two. Preferences are also usually assumed to be transitive, so that if x is
preferred to y and y is preferred to z, then x is also preferred to z. Where uncertainty is involved,
preferences are usually assumed to satisfy an “independence” or “cancellation” or “sure-thing”
condition (e.g., Savage’s axiom P2), which ensures that comparisons among acts depend only on
the events where they lead to different consequences.3 For example, if x and y are two acts that
yield different consequences if event E occurs but yield the same consequence if E does not
occur, and if two other acts x and y agree with x and y, respectively, if E occurs but yield a
different, though still common, consequence if E does not occur, then x is preferred to y if and
only if x is preferred to y. This implies that numerical representations of preference can be
additively decomposed across events, setting the stage for expected-utility calculations.
Preferences are also assumed to satisfy conditions ensuring that the effects of beliefs about events
can be separated from the effects of values for consequences. One such condition (Savage’s
axiom P3) requires that “value can be purged of belief.”4 Suppose that acts x and y yield
identical consequences everywhere except in event E, where x yields consequence c1 and y yields
c2. Then a preference for x over y suggests that consequence c1 is more highly valued than
consequence c2. Now consider two other acts x and y that yield identical consequences
everywhere except in event E, where they lead to c1 and c2 respectively. (There is an additional
technical requirement that the events E and E should be “non-null”—i.e., not regarded as
impossible.) Then x is assumed to be preferred to y if and only if x is preferred to y, which
means the ordering of preference between two consequences cannot depend on the event in
which they are received. Another such condition is that “belief may be discovered from
preference” (Savage’s axiom P4). Suppose that consequence c1 is preferred (as a sure thing) to
consequence c2, and suppose that for two events A and B, the lottery in which c1 is received
conditional on A and c2 is received conditional on not-A is preferred to the lottery in which c1 is
received conditional on B and c2 is received conditional on not-B. (This suggests that A is
regarded as more probable than B.) Then the same direction of preference must hold when c1 and
c2 are replaced by any two other consequences c1 and c2, respectively, such that c1 is preferred
to c2. A further condition that is needed to uniquely separate belief from value, but which is
3
The independence axiom has been the object of much controversy, and relaxations of it have led to many varieties
of “non-expected utility” theory—e.g. Machina (1982), Chew (1983), Quiggin (1982), Yaari (1987), Schmeidler
(1989).
These labels for Savage’s axioms P3 and P4 are due to Shafer (1986), who points out that they are actually more
important than the more famous independence axiom because of the strain they place on the notion of a “small
world.”
4
5
often left implicit, is that the strength of preference between any two given consequences—e.g.,
the perceived difference between “best” and “worst”—must have the same magnitude and not
merely the same sign in every state of nature (Schervish et al. 1990).
The axioms imposed on preferences yield a representation theorem stating that to every event not
under her control the agent can assign a numerical degree of belief called a probability, and to
every consequence she can assign a numerical degree of value called a utility, according to which
the more preferred of two acts is the one that yields the higher expected utility. (A particularly
elegant axiomatization of subjective expected utility is given by Wakker 1989.) To continue our
earlier example, this means that descriptions of states can be summarized by their corresponding
probabilities, and descriptions of consequences can be summarized by their corresponding
utilities in tables of acts and alternatives, as follows.
Table 3:
Alternative:
Walk to work with umbrella (a1)
Walk to work without umbrella (a2)
Probabilities and utilities:
p1
p2
u11
u12
u21
u22
Here p1 and p2 denote the agent’s probabilities of “rain” and “no rain,” respectively, u11 denotes
the utility level of the consequence “get slightly wet, hassle”, and so on. The expected utility of
alternative i is then given by:
EU(ai) = p1ui1 + p2ui2,
and she strictly prefers a1 over a2 if and only if EU(a1) > EU(a2).
Thus, the rational individual is a maximizer, and the parameters of the objective function she
maximizes are her beliefs and values, represented by numerical probabilities and utilities.
Beliefs and values are imagined to have different subjective sources—the former depending on
information and the latter on tastes—and their effects are imagined to be separable from each
other. Only beliefs (probabilities) matter in problems of reasoning from evidence, while only
values (utilities) matter in problems where there is no uncertainty.
One thing remains to complete the description of individual rationality: a link between mind and
body, or between preference and choice. This is the axiom of rational choice, which states that
the event that occurs shall be one in which every agent chooses her most-preferred alternative.5
To continue the example above, if the agent’s preferences are such that EU(a1) > EU(a2), then
5
The axiom of rational choice is not always listed as a formal assumption. Sometimes it is merely remarked that
“choices follow preferences” or “preferences are revealed by choices,” or else it is assumed that preference and
choice are synonymous: the agent’s “choice function” is simply given in lieu of a preference ordering. The rational
choice axiom does explicitly appear in Luce and Raiffa (1957) as the “law of behavior”: assumption ix in Chapter 3,
and it is informally used by Savage. Here we maintain a careful distinction between “preference” and “choice” to
call attention to the question of which choices are actually observed and which are not: a “preference” is a choice
that may be (and usually is) merely hypothetical because the alternatives in question are never actually faced together
(and are often nonexistent or even logically impossible).
6
the theory predicts that event a1 will occur: she will carry the umbrella. The axiom of rational
choice thereby gives operational meaning to the concept of preference. For, a preference for aj
over ak is meaningless by itself if aj and ak are merely hypothetical acts rather than feasible
objects of choice. But through the axioms imposed on individual preferences, such a preference
constrains the preferences that may exist with respect to other acts, say a1 and a2, that are
feasible, and through the axiom of rational choice it then exerts an indirect effect on choice. In
the standard theory, most preferences have only this tenuous connection with materially
significant behavior.
b. Strategic rationality: The axioms of individual rationality provide a solution to the “one
body problem” of rational choice, reducing it to a numerical maximization exercise. The
solution to the “problem of 2, 3, 4,... bodies” seemingly requires more powerful tools, as noted
by von Neumann and Morgenstern (1944, p. 11):
“Thus each participant attempts to maximize a function... of which he does not control all
variables. This is certainly no maximum problem but a disconcerting mixture of several
conflicting maximum problems. Every participant is guided by another principle and
neither determines all variables which affect his interest. This kind of problem is nowhere
dealt with in classical mathematics.” 6
Not only are the participants guided by different principles, but their beliefs about events not
under their own control are also informed by the knowledge that some of those events are
controlled by other rational individuals like themselves. A decision problem in which two or
more agents make choices that affect each other’s interests is called a “game of strategy,” and the
theory of such games tries to account for the simultaneous optimization of different objectives
and the phenomenon of reciprocal expectations of rationality. It does so by extending the
assumptions of individual rationality with assumptions of strategic rationality.
First, it is explicitly assumed that individual rationality is common knowledge, so that every
player in the game knows that every other player makes choices so as to maximize expected
utility, according to his or her own probabilities and utilities. The assumption of common
knowledge of rationality constrains the probabilities that one agent may assign to the choices of
another. For example, if Alice knows that Bob’s alternative x is dominated by his alternative y—
i.e., if the consequences of y for Bob are strictly preferred by him to the consequences of x in all
outcomes of events—then she must assign zero probability to the event that Bob will choose x
when y is available. Furthermore, Bob knows that Alice knows this—which constrains the
beliefs that Bob may hold about Alice’s beliefs—and Alice knows that Bob knows she knows it,
and so on in an infinite regress.
In order for the infinite regress of beliefs to get started, the players must first know something
definite about the structure of each other’s decision problems—e.g., Alice must know something
about Bob’s consequences and about his preferences among those consequences. It is therefore
normally assumed that utility functions are common knowledge, which implicitly requires not
6
We will show later that the problem of one body and the problem of two or more bodies can be solved with the
same mathematics after all: there is really no essential difference between individual and strategic rationality.
7
only that Alice knows Bob’s mapping of events to consequences, but also that she knows Bob’s
preference ordering over all acts that can be constructed from those events and consequences.
Taken literally, that is a tall order. For purposes of modeling, it is usually just assumed that the
numerical utility of every outcome of events for every player is commonly known—or that the
players will somehow act as if this is so. Such information about utilities is summarized in a
payoff matrix that constitutes the “rules of the game.” For example, the payoff matrix for a 2player game might have the following form:
Table 4
Top (a11)
Bottom (a12)
Left
(a21)
1, -1
-1, 1
Right
(a22)
-1, 1
1, -1
Here, player 1 (“row”) has two alternatives, Top (a11) and Bottom (a12), while player 2
(“column”) has two alternatives, Left (a21) and Right (a22). The numbers in each cell are the
consequences for the respective players measured in units of personal utility. This particular
game is known as “matching pennies.” The story line is that each player has a penny and they
simultaneously choose to show heads or tails. Player 1 wins her opponent’s penny if both show
heads or both show tails, and player 2 wins her opponent’s penny if they show heads-tails or
tails-heads. It is assumed that the players have constant marginal utility for money and that no
other stakes are riding on the game, hence their utility payoffs are 1.
The requirement that players know each other’s utility functions is often relaxed by permitting
the agents to be uncertain about each other’s utilities, which gives rise to a game of incomplete
information. The introduction of incomplete information greatly extends the range of strategic
situations that can be modeled by game theory—e.g., situations such as auctions, in which the
players’ uncertainty about each others’ values is pivotal. But the increased generality of
incomplete-information games comes at a price, namely that additional strong assumptions are
needed to keep the uncertainty models tractable. First, it must be assumed that the set of possible
utility functions for each player can be reduced to a manageable number of “types,” which are
themselves common knowledge. Second, it is necessary to constrain the reciprocal beliefs that
may exist with regard to such types. For example, if Alice and Bob are uncertain of each other’s
types, Alice’s beliefs about what Bob believes about her type, given his own type, should be
consistent with what Bob really believes, and so on in another infinite regress. This sort of
consistency is usually enforced via the common prior assumption (CPA) introduced by Harsanyi
(1967). The CPA states that there is a commonly-held prior distribution over types from which
the actual probability distribution of each player concerning all the others’ types can be derived
by conditioning on her own type.
The following is an example of an incomplete information game in which player 1 is uncertain
about the type of player 2, which may be A or B (Myerson 1985, Nau 1992b). As in the previous
example, player 1 chooses between Top and Bottom and player 2 (whatever her type) chooses
between Left and Right. However, the utility payoffs to both players now also depend on the
type of player 2. It is assumed that there is a common prior distribution assigning probabilities of
8
60% and 40%, respectively, to types A and B of player 2. These are the probabilities that player 1
assigns to player 2’s type, and player 2 knows that player 1 assigns, and player 1 knows that
player 2 knows that player 1 assigns, and so on. Meanwhile, player 2 knows her own type with
certainty at the instant she makes her move, and player 1 knows that she knows it, and so on.
Table 5
Top (a11)
Bottom (a12)
Type A (60%)
Left (a21) Right (a22)
1, 2
0, 1
0, 4
1, 3
Top (a11)
Bottom (a12)
Type B (40%)
Left (a21)
Right (a22)
1, 3
0, 4
0, 1
1, 2
The common prior assumption is not only central to the theory of incomplete-information games,
but it is also implicitly invoked in the solution of complete-information games when randomized
choices are used. The randomizing probabilities are then assumed to be common knowledge,
which is equivalent to assuming a common prior distribution on outcomes of the game. The
CPA is widely used elsewhere in information economics—e.g., in models of trading in markets
under uncertainty (Milgrom and Stokey 1981)—and it is also sometimes invoked in discussions
of social justice. (See Harsanyi 19xx, Heap et al. 1992, Binmore 1994 for discussions of the
latter.)
A further assumption that is normally made in games involving randomized choices and/or
incomplete information is that the probability distributions held by players with respect to their
opponent’s types and choices are independent of their own choices given their own types. In
other words, the players are assumed to act as if they do not believe it is possible to correlate
their choices with the simultaneous choices of their opponents. A rationale often given for the
independence assumption is that it confers generality by not requiring the existence of
correlation devices. The independence assumption also seems consistent with the spirit of
“noncooperative” play from the perspective of an individual player: it suggests that she has
complete freedom to do whatever she wants at the instant she moves her hand. Last but by no
means least, the independence assumption is mathematically convenient: it yields tighter
predictions than would be obtained if correlated beliefs were allowed.
The assumptions of common knowledge of rationality, common knowledge of utilities, common
prior probabilities, and independence imply that the outcome of a game should be an equilibrium
in which every player’s choice is expected-utility-maximizing given her own type and in which
the players’ reciprocal beliefs are mutually consistent and conditionally independent of their own
choices given their own types. In a game of complete information (known utility functions),
such an equilibrium is a Nash equilibrium (Nash 1951), and in a game of incomplete information
it is a Bayesian equilibrium (Harsanyi 1967).
Examples, continued: The game of Table 4 has a unique Nash equilibrium in which
each player chooses randomly between her two alternatives with equal probability—i.e.,
she flips the coin to decide whether to show heads or tails. The game of Table 5 has a
9
unique Bayesian equilibrium in which player 1 always plays Top and player 2 plays Left
if she is type A and Right if she is type B.7
Nash and Bayesian equilibrium are the most commonly used solution concepts for
noncooperative games, but they do not always (or even generally) yield unique or intuitively
reasonable solutions, and so other assumptions are often invoked to “refine” the set of equilibria.
For example, it may also be required that the equilibrium be symmetric, robust against errors in
play or the occurrence of zero-probability events, stable against defections by more than one
player, and so on.
c. Competitive rationality: When agents interact in competitive markets, the situation is
analogous to a large ensemble of particles in physics, and the convoluted mechanics of games of
strategy can be replaced by a simpler statistical mechanics. Agents don’t need to know the
details of every other agents’ preferences and strategies: each is affected only by the aggregate
behavior of the others, and interactions among them are typically mediated by an impersonal
mechanism of prices. The relevant “consequences” for market participants usually are vectors of
commodities (consumption goods) that can be exchanged or produced. Every agent has an
initial endowment of commodities and, under a given schedule of prices, she can exchange a
bundle of commodities for any other bundle of equal or lesser cost—assuming that buyers and
sellers can be found at those prices. Or, if she is a producer, she can convert a bundle of
commodities into a different bundle through the use of technology. The agent’s alternative set
consists of exchanges and/or production plans that are feasible in the sense of lying within her
budget and/or technology constraints at the prevailing prices.
Two assumptions of competitive rationality are commonly made in models of markets. First, it is
assumed that all agents are “price takers,” meaning that they choose alternatives to maximize
their own expected utilities at the given prices; they do not attempt to manipulate those prices.
Second, it is assumed that prices are determined so that all markets clear—i.e., so that the
optimal choices of all agents are jointly as well as individually feasible—and that this somehow
occurs prior to any exchange or production. All transactions are then made at the marketclearing prices. The latter assumption is a kind of common knowledge assumption with respect
to aggregate preferences: the market-clearing prices are “sufficient statistics” for the preferences
of all agents. (The mechanism by which prices become known usually is not modeled, although
suggestive tales are told of a “Walrasian auctioneer” or central pricing authority to whom agents
truthfully report their preferences and who then computes and posts the market-clearing prices.)
When the assumptions of price-taking and market-clearing are satisfied, a general competitive
(Walrasian) equilibrium is said to exist.
d. Social rationality: Social rationality is rationality as seen from the viewpoint of an
omniscient social planner. The principal requirement of social rationality, which is central to
7
Note that Left is a dominant—i.e., always better—strategy for type A of player 2 and Right is a dominant strategy
for type B. If player 2’s type is revealed to player 1 before both players move, then player 1 will play Top when 2 is
type A (correctly anticipating that 2 will play Left) and 1 will play Bottom if 2 is type B (correctly anticipating a
move of Right). In the incomplete-information version, player 2 benefits from player 1’s uncertainty and is able to
obtain a higher payoff when her type is B.
10
social choice theory as well as market theory, is that the final endowments of the agents should
be Pareto efficient: it should not be possible to redistribute resources in such a way that
everyone is at least as well off (according to her own preferences) and at least one agent is strictly
better off. The First and Second Theorems of welfare economics establish a duality relationship
between social and competitive rationality, namely that a competitive equilibrium is Pareto
efficient, and conversely, under suitable restrictions on preferences, every Pareto efficient final
endowment is a competitive equilibrium with respect to some initial endowment (Arrow 1951a).
e. Rational asset pricing: In the literature of financial economics, the rationality of asset prices
is defined by the requirement of no arbitrage, namely that there should be no riskless profit
opportunities (money pumps or free lunches) . Many important results in the theory of finance—
formulas for pricing derivative securities (Black and Scholes 1973, Merton 1973) and valuing
cash flow streams (Ross 1978), the capital structure theorem (Modigliani and Miller 1958), and
factor models of market risk (Ross 1976ab)—are based on arbitrage arguments rather than on
detailed models of individual preferences. Indeed, such results are sometimes characterized as
“preference free,” although aggregate preferences presumably determine the prices of the
fundamental assets from which other prices are derived by arbitrage. (For broader perspectives
on arbitrage, see Garman 1979, Varian 1987, Ross 1987.) No-arbitrage is closely related to
Pareto efficiency: a commonly known Pareto inefficiency would constitute an arbitrage
opportunity for a middleman if transaction costs were negligible and agents behaved
competitively rather than strategically. The no-arbitrage requirement is also known as the
“Dutch book argument” in the theory of gambling: the odds posted by a rational bookmaker
should not present the opportunity for a sure win (“Dutch book”) to a clever bettor.
f. Rational learning: In models where events unfold gradually over time, it is necessary to
describe how agents learn from their experiences. The learning process is usually modeled via
Bayes’ theorem, in which “prior” probabilities are converted to “posterior” probabilities through
multiplication by a likelihood function that summarizes the incoming data. This kind of
“Bayesian learning” appears in models of intelligent search, intertemporal asset pricing,
information acquisition, sequential games, entrepreneurship, and research and development.
g. Rational expectations: In models of multi-period markets, an additional assumption of
rational expectations is often imposed, namely that agents form beliefs about future economic
events that are self-fulfilling when rationally acted upon. Thus, the agents’ probabilities are not
only mutually consistent but objectively correct, as if the agents all understand exactly how the
economy works, making them not only as smart as the theorist but perhaps even smarter.
Although this assumption lends new meaning to heroism, it nevertheless captures the intuitive
idea that today’s values of economic variables such as prices and interest rates are determined
partly by agents’ expectations of their future values as well as by their expectations for
underlying fundamental variables and their preferences for intertemporal substitution.
Summary: The preceding outline of rational choice theory is admittedly over-simplified: it is a
deliberate attempt to compress the theory into a tidy logical structure, when in fact it is a rather
bulky grab-bag of tools and methods used in different configurations by different academic
tribes. Some tribes mainly study individual decisions, others study markets, others focus on
11
games of strategy, others study problems of collective action, and so on. The field is further
subdivided into those who model decisions only under conditions of certainty (spiritual
descendants of Walras, Edgeworth, and Pareto), those who model uncertainty in terms of
objectively given probabilities or relative frequencies (descendants of von Neumann and
Morgenstern) and those who regard probabilities as being subjectively determined (descendants
of de Finetti and Savage). Nevertheless, certain generalizations are valid. A typical rational
choice model of a social or economic situation begins with a description of actors, events, and
consequences. The actors’ beliefs about events and values for consequences are then quantified
in terms of appropriate probabilities and utilities and—where multiple actors are involved—a
suitably refined equilibrium concept is invoked. A solution (preferably unique) is exhibited in
which all agents simultaneously maximize their expected utilities. The analysis concludes with a
study of comparative statics: how would the equilibrium solution be affected by changes in the
physical or psychological initial conditions? And within this framework, the role of institutions
(households, firms, markets, governments, etc.) is to enable the agents to increase their expected
utilities, according to the probabilities and utilities with which they were originally endowed, by
choosing wisely from among the alternatives they were originally given.
2.2 Trouble in paradise
As measured by pages of journal output, positions held in leading university departments, and
academic prizes won, the rational choice paradigm has been hugely successful. It has aspired to,
and some would argue it has achieved, the status of a universal theory of social and economic
behavior, comparable in elegance and scope to grand theories of the physical sciences. But by
other measures its record has been spotty—some would argue even dismal—and the bashing of
rational choice theory (and bashing back by its supporters) has become an increasingly popular
pastime even as the theory has extended its reach into discipline after discipline. See, for
example, the edited volumes by Barry and Hardin (1982), Hogarth and Reder (1985), Elster
(1986), Moser (1990), and Cook and Levi (1990) which present contrasting perspectives from
economists, psychologists, and philosophers; Arrow’s (1987, 1994) articles on rationality and
social knowledge; Sugden’s (1991) survey of foundational problems; the critical survey by Heap
et al. (1992); the special issue of Rationality and Society (1992) devoted to a symposium on
rational choice; Green and Shapiro’s (1994) provocative critique of rational choice in political
science and the lively volume of rebuttals and rejoinders edited by Friedman (1995). Here I will
focus on a few of the problems with rational choice theory that are especially troubling from a
foundational perspective (and which, not incidentally, I will aim to fix in the sequel).
Consequences and acts: It is questionable whether some of the primitive elements of rational
choice theory hold water even as idealizations of intelligent and self-interested behavior. The
very definition of terms is problematic: what exactly is a “consequence” and how detailed must
its description be? At what point in time and at what level of perception is it experienced? For
example, “get drenched” is rather vague as a description of a consequence, because dousing with
water might be enjoyable in some cases and not in others. So, perhaps the description should
include qualifiers such as the ambient temperature, the social setting, one’s state of dress or
undress, the element of surprise, and so on. But then it becomes hard to separate the definition of
12
a consequence from the definition of an act: is “get drenched by cold rain while walking to work
in a three-piece wool suit” a consequence (a result that can be obtained regardless of the state of
the weather or your selection of attire) or an act (an assignment of sensory experiences to
events)? The ramifications of getting caught in the rain are not all immediate or certain, either—
you might then catch pneumonia or be embarrassed at an important meeting or incur the cost of a
new suit, with further repercussions for health, employment, consumption, etc. Savage (1954)
recognized these difficulties and introduced a distinction between “small worlds” and the “grand
world,” the former being what we model and the latter what we actually inhabit. A consequence
in the small world then corresponds to an act in the large world. But this philosophical trick does
not make the practical difficulty go away, because consequences as modeled still need to satisfy
assumptions that may be implausible in a small-world setting. A consequence (such as“get
drenched by cold rain”) needs to be assignable to any state of nature (including “no rain”) and
have a utility that is independent of the state. (See the marvelous exchange of letters between
Aumann and Savage on this point, which is reprinted in Drèze 1987. A deep analysis of the
problem of separating states-of-the-person from states-of-nature is given by Shafer 1986.) We
are on somewhat firmer ground in situations where the consequences are tangible commodities—
as in models of pure exchange—but there too, where uncertainty is involved, it is problematic to
assume that preferences for consequences are independent of the states in which they are
obtained.
Preferences: The concept of “preference” is so deeply embedded in modern choice theory that it
is nearly impossible to imagine life without it. The marriage between the familiar linguistic
concept of a preference between two alternatives and the topological concept of an ordering
relation seems just too convenient to pass up. Yet there are grounds for questioning whether
preference is a suitable behavioral primitive for a theory of choice—particularly a theory that
aspires to make quantitative predictions and/or to encompass more than one agent at a time.
Indeed, the reification of individual preferences is arguably the root of most of the problems and
paradoxes of standard rational choice theory.
First, the concept of a preference between x and y presupposes that the alternatives x and y are
already in some sense “given,” whereas the most difficult problem in everyday decision-making
is often the imaginative process of creating or discovering the alternatives. To say that
rationality consists of having preferences that are sufficiently well-behaved is to ignore the very
large problem of where the alternatives come from. (Further discussion of this large problem
will be put off until section 4.)
Second, even in situations where the alternatives are already given, the articulation of preferences
is neither necessary nor always helpful as an antecedent to choice. Granted, in problems of
discrete choice we often try to rank-order the alternatives or make pairwise comparisons, but we
do not routinely compare large numbers of purely imaginary or even counterfactual alternatives.
In many real decision problems—especially problems where quantitative precision is
important—we think in direct numerical terms such as “reservation prices” or “willingness to
pay” for which there are convenient reference points in the marketplace. In decision analysis
applications, numerical assessments of subjective probabilities and utilities are typically carried
out not by asking for endless binary comparisons, but by asking subjects to solve matching
13
problems in which a price or a lottery probability is adjusted until it becomes equivalent to some
real prospect. Shafer (1986) argues that preferences do not even exist as latent mental states—
rather, they are constructed only when the need arises by reasoning from evidence and by
framing the comparison in a manner that evokes an appropriate set of values or goals. This
constructive view of preference is widely shared by behavioral researchers and decision analysts.
Keeney (1992) shows that the process of articulating and structuring one’s values, en route to the
creation of alternatives and the determination of preferences among them, can be quite complex.
Third, even where preferences already exist or have been constructed, they may hard for others to
observe. Preferences are a form of private, mental behavior that may precede or cause physical
behavior. But for preferences to be a useful construct in a theory of multiple agents, it is
necessary for agents to know something about each others’ preferences independently of their
choices. How does this happen? Why would agents wish to reveal their preferences to each
other, and through what communication mechanism could they credibly do it prior to making
choices? The problem of communicating preferences is not often modeled—rather, agents are
imagined to discern each other’s preferences through intuition and experience—but when it is
modeled, rather serious difficulties of “incentive compatibility” emerge (Groves and Ledyard
1988).
The problems of defining consequences and observing preferences obviously have profound
implications for assumptions about strategic and competitive rationality. If agents cannot
vicariously perceive each others’ consequences and their preferences among all acts composed of
those consequences, the assumptions that utility functions are common knowledge or that
market-clearing prices are revealed prior to trade appear dubious.
The axioms: The best-known complaints against rational choice theory are empirical
demonstrations that individuals often violate the preference axioms, even in laboratory
demonstrations or thought experiments, and even when the subjects are expert decision makers
or eminent decision theorists. Rather, they display an assortment of heuristics, biases, and errors
in mental accounting: they exhibit reference-point effects, they evaluate gains and losses
asymmetrically, they are sensitive to the language in which decision problems are framed, their
preferences do not display the “linearity in probabilities” that would characterize expected-utility
maximizers (particularly when probabilities go to zero or one), and they are sometimes willfully
intransitive. In response to empirical failures of the standard theory, much of the foundational
work on rational choice over the last 20 years has focused on relaxations of the axioms of
independence, transitivity, continuity, etc. These axiomatic explorations have yielded elegant
new theories of “non-expected” utility, such as generalized expected utility, rank-dependent
utility, skew-symmetric-bilinear utility, lexicographic utility, non-additive probability, etc., which
are able to account for some of the behavioral anomalies. (See Machina 1987, Fishburn 1988,
Chew and Epstein 1989, Chew et al. 1993, for surveys of such models) Insofar as the primitives
of the theory—consequences, acts, and preferences—may be ill-defined or unobservable, it is
possible that all this axiom-tweaking has yet failed to address the most serious problems, a point
to which we shall return. Nevertheless, if complaints are to be lodged against specific axioms, it
would perhaps be better to single out the axiom of completeness and the axioms which enforce
the separation of probability and utility.
14
The completeness axiom requires that the agent be able to assert a direction of preference (or else
perfect indifference) between any two acts, which is objectionable for several reasons. First, it
places an unreasonable cognitive burden on the agent: she is required to have a preference
between any two acts that can be constructed from the given sets of events and consequences, no
matter how far from everyday experience and no matter how fine the distinctions between them.
Second, it increases the strain on all the other axioms: if the agent were allowed to say “I don’t
know” in response to some questions about preferences, she would be less likely to violate, say,
transitivity or the sure-thing principle. Indeed, systematic violations of the preference axioms by
experimental subjects tend to occur in situations cunningly designed to straddle indifference
points. Models of subjective probability and expected utility without the completeness axiom
have a long history in the literature (e.g., Koopman 1940, Smith 1961, Good 1962, Aumann
1962, Levi 1980, Walley 1991, Nau 1992, Seidenfeld et al. 1998). Schmeidler (1989) remarks
that “the completeness of the preferences seems to me the most restrictive and most imposing
assumption of the theory.” Nevertheless, models of incomplete preferences have remained
largely outside the mainstream, perhaps because of the mathematical awkwardness of
representations that are expressed in terms convex sets of probabilities and utilities rather than
point values, and perhaps because of the theorist’s natural desire for a theory that leaves no
things undecided. Even Schmeidler did not abandon the completeness axiom, but merely
shortened its reach through an assault on the independence axiom.
The axioms which ensure that value can be purged of belief and that belief can be discovered
from preference, and the further implicit assumption that strengths of preference between
consequences are invariant across states, are needed in order to uniquely separate probabilities
from utilities. As such, they embody a traditional normative view that beliefs and values have
different subjective sources: beliefs ought to be based on information whereas values ought to be
based on wants and needs (and occasionally on ethical principles). A person who lets her values
contaminate her beliefs is derided as a wishful thinker, while one who lets her beliefs
contaminate her values is lacking in moral fiber. And certainly the mental exercise of separating
beliefs from values is an aid to clear thinking in the practice of decision analysis. Yet, as Shafer
(1986) argues, choice is often goal-oriented in practice, and the value attached to an outcome
may depend on its feasibility of attainment if it suffices to meet an important goal. (This is one
possible explanation of the Allais paradox: the smaller jackpot may be perceived as more
attractive when it is a sure thing than when it is just another long shot.) But regardless of
whether an agent can privately separate her own beliefs from her own values, the relevant
question in a multi-agent problem is whether one agent can separate the beliefs and values of
another. If one agent cannot look inside another’s head, but instead can only observe the choices
she makes among feasible alternatives with material consequences, then the effects of probability
are confounded with the effects of unknown prior stakes in the outcomes of events and/or
intrinsic state-dependence of utilities for consequences (Kadane and Winkler 1988; Schervish,
Seidenfeld, and Kadane 1990; Nau 1995b). And if it is difficult for agents to separate each
others’ probabilities from their utilities, then the traditional game-theoretic assumptions of
commonly held prior probabilities and commonly known utilities are again called into question.
15
An even deeper question is whether the usual axioms of subjective probability are valid in the
context of a noncooperative game. Savage’s axioms draw an important distinction between the
decision maker’s choice among alternatives and nature’s choice among the states—and the latter
is required to be independent of the former. When these axioms are invoked to justify the use of
subjective probabilities to represent a player’s uncertainty about the outcome of a noncooperative
game, it is necessary to treat the choices of the other players as if they were exogenous states of
nature. But the player’s reasoning about her own choice may endogenously affect her
expectations about the choices of others, an issue which has been raised by Sugden (1991) and
Hammond (1997).
Equilibrium: The solution to an interactive decision problem in the standard theory is typically
expressed as an equilibrium in which every agent chooses a strategy (which may be a single
alternative or a probability distribution over alternatives), and the strategy of every agent is an
optimal response to the strategy of every other agent. Furthermore, the analysis often involves an
implicit leap from an initial state of disequilibrium to a final state of equilibrium. The agents
arrive at the marketplace with their initial endowments, market-clearing prices are determined,
then everyone trades and goes home. Or agents open the game box, read the rules, then
independently reason their way to commonly known equilibrium strategies. The subsequent
comparative statics exercise asks questions of the form: how would the final solution vary with
the initial conditions? On the surface, the before-to-after leap appears to lie within the proper
scope of the theory: the expected-utility model of the rational agent predicts not only what she
will do in the given situation, but also what she might do in any imaginable situation.
Yet the initial state of disequilibrium is usually fictitious and the dynamic principles that might
propel agents toward a state of equilibrium are not specified: the theory describes equilibrium
behavior, not disequilibrium behavior. In recent years, considerable attention has been focused
on theoretical and experimental models of convergence to equilibrium via learning, adaptation,
and natural selection, usually by iterating the market or game situation a large number of times.
But many interesting decision problems are intrinsically novel or unique—e.g., what will happen
when a new product is launched or a new market is opened or a new competitor enters the game?
Furthermore, the most popular solution concepts for noncooperative games (Nash and Bayesian
equilibrium) are not guaranteed to yield unique equilibria, and in cases where the equilibrium is
not unique, further restrictions (e.g., symmetry, perfection, etc.) are often imposed somewhat
arbitrarily. Similarly, the uniqueness of a competitive equilibrium depends on the high-handed
assumption that no trade takes place out of equilibrium. More realistically, the outcome of
exchange among agents will be one of the many Pareto-optimal points on the “contract curve,” as
recognized by Edgeworth (1881). The lack of dynamic principles and the non-uniqueness of
equilibria leave us on shaky ground in asserting that the same generic solution will obtain if the
initial conditions of a model are perturbed in a comparative statics analysis.
Impossibility: Despite its aspirations of explaining the behavior of agents in any social
configuration, the theory often finds that it is hard, if not impossible for rational agents to make
decisions as a group. Arrow’s (1951b) famous theorem shows that there is no way (other than
dictatorship) to aggregate the ordinal preferences of a group in such a way that the group’s
preferences are rational and at the same time respectful of consensus. Similar results apply to
16
voting systems (they are subject to manipulation—Gibbard 1973 and Satterthwaite 1975) and to
the aggregation of probability judgments by different experts (there is no combining rule that is
consistent with respect to the marginalization of probabilities and also “externally Bayesian”—
Genest & Zidek 1986). When values are represented by cardinal utility functions, which are
meaningful under conditions of risk or uncertainty, a Pareto optimal allocation can be found by
solving a social planner’s problem in which a weighted sum of individual utilities—a sort of
group utility function—is maximized. But then the agents have incentives to conceal or
misrepresent their true utilities, and even if their true utilities could be extracted, the choice of
weights would remain problematic because of the incomparability of utilities between
individuals. These results suggest that there is no natural way to define the collective beliefs and
values of a group of rational agents, nor is there any natural institutional mechanism for reaching
a consensus. Those who know their own minds too well have little use for each others’ beliefs
and values except as instruments for increasing their own expected utilities, and collective action
is hard to organize.
Simple questions, equivocal answers: Even apart from the realism of its assumptions about
individual rationality or the nature of equilibria, the empirical success of rational choice theory in
explaining existing institutions has been controversial. Simple questions turn out to not have
simple answers, even in the field of economics where rational choice models originated and
where beliefs and values are most readily quantified. For example, what is the mechanism of
price formation in competitive markets? (A variety of mechanisms are observed in practice, and
they are often highly decentralized or even chaotic.) Why do agents trade vigorously on stock
exchanges in response to releases of information? (If everyone is rational with common prior
probabilities, no one can profit from the receipt of private information. It is necessary to
introduce measured amounts of irrationality into a securities market—such as the presence of
“noise traders”—to explain how rational traders might be able to benefit from private
information.) Why are the real rates of return on government securities so low relative to rates
of return on stocks? (The implied coefficient of risk aversion for a typical investor is
pathological, a finding known as the “equity premium puzzle.”) Why do firms exist, and does a
firm have a utility function? (In the classical theory of the firm, it is treated as a unitary actor
whose utility function is profit. In game-theoretic models of duopoly and oligopoly, the firm
maximizes a utility function that may be profit but may also have other attributes. In the theory
of finance, the firm seeks to maximize its value—that is, the expected net present value of its
future cash flows—but it has no particular corporate identity: its managers merely serve the
interests of the shareholders. If those shareholders are well-diversified, they will be unconcerned
about the possibility of financial distress or bankruptcy and will wish the firm to act in a riskneutral fashion on their behalf. In practice, firms do have somewhat distinct identities and
corporate cultures, they are constrained by internal and external regulations that sometimes
inhibit and sometimes promote risk-taking, and their managers often appear to be concerned
about the interests of other stakeholders such as themselves.)
When we turn to non-economic disciplines such as political science, the empirical success of
rational choice theory is even harder to validate. Here too, the theory has difficulty explaining
simple facts, such as why the citizens of a democracy would bother to vote in substantial
numbers. (The expected impact of one’s own vote on the outcome of an election is usually nil,
17
while the cost of voting is finite, hence individuals should not bother to vote, a result known as
the “voter’s paradox.” In order to explain voter turnout on the order of 50% in state or national
elections, it is necessary to imagine that voters derive utility from conforming with social norms,
which runs counter to the egoistic image of the utility maximizer.) In a sweeping critique of
rational choice models in American politics, Green and Shapiro (1994) conclude:
...the case has yet to be made that [rational choice] models have advanced our
understanding of how politics works in the real world. To date, a large proportion of the
theoretical conjectures of rational choice theorists have not been tested empirically.
Those tests that have been undertaken either have failed on their own terms or garnered
theoretical support for propositions that, on reflection, can only be characterized as banal:
they do little more than restate existing knowledge in rational choice terminology.
The fossil record: Another interesting body of empirical evidence is the fossil record of rational
choice applications in the practice of management. Decision tree analysis and game theory have
been taught in business schools and economics departments for almost forty years, but they have
had only a modest impact—if any at all—on the decision-making practices of most firms and
individuals, and today they are being elbowed out of the curriculum by non-quantitative
“strategy” courses. There is as yet little proof of the superiority of rational-choice-based decision
models over competing models that do not emphasize expected-utility maximization or
equilibrium arguments. (Methods of uncertainty modeling and decision analysis that expressly
violate the assumptions of rational choice theory—such as frequentist statistics and the analytic
hierarchy process—have yet to be exterminated by natural selection.) Successes have been
achieved in some industries and application areas—e.g., medical decision making, oil and gas
exploration, pharmaceutical research and development, environmental modeling, auction
design—but those stand out as the exceptions that prove the rule. Methods for making “optimal”
decisions are surprisingly hard to sell and hard to implement given their importance in theory.8
And where formal methods of decision analysis and game theory are successfully applied, their
main benefit rarely lies in finding numerically optimal solutions to well-specified problems.
Rather, the benefit of formal methods more often lies in their potential for helping participants to
frame the issues, communicate with other stakeholders, discover creative solutions, and
manipulate the rules of the game—activities that take place outside the boundaries of the theory.
Summary: The litany of problems cited above is not intended to be exhaustive or even-handed.
Rather, it is a helicopter tour of some significant unresolved foundational issues and empirical
anomalies. Proponents of rational choice theory have familiar disclaimers: expected utility
theory does not describe how people actually behave, rather it is a normative theory that ought to
help them behave better. Game theory does not predict the result of strategic interactions; rather,
it is only a language for discussing the forms that such interactions might take. Economic theory
doesn’t really expect consumers and investors to understand the intricacies of the market; that
assumption is merely a device for generating models with testable implications. Agents are not
self-conscious maximizers; they only act “as if” they were—evolution and learning will see to
8
Models of rational asset pricing have been enormously successful in finance, but the most robust and durable asset
pricing models rely on arbitrage arguments and statistical regularities rather than detailed models of behavior at the
agent level.
18
that. But the following rejoinders apply. First, real individuals do not come close to satisfying
the assumptions of standard rational choice theory, nor do they find the theory to be particularly
helpful in their everyday decision-making. Second, if they did satisfy the standard assumptions,
they would actually find it rather pointless to conduct business and politics as we recognize them.
Finally, the devil is in the details: rational choice theory does not predict much of anything
unless it is wielded by skilled hands and embellished with well-calibrated probabilities and
utilities, heroic assumptions of common knowledge and congruent beliefs, and ad hoc restrictions
on the nature of equilibria.
2.3 Alternative paradigms
Although rational choice theory is the dominant paradigm of quantitative research in the social
and economic sciences, it is not the only such paradigm—and of course not all social-scientific
research is quantitative. The following is a brief survey of alternative paradigms that have
useful lessons to offer.
Bounded rationality, behavioral economics, and organization theory: Over the same period
in which rational choice theory has flowered, a parallel theory of bounded rationality in
economics and organization theory has been developed by Herbert Simon, James March, Richard
Cyert, Sidney Winter, Richard Nelson, Richard Thaler, and many others. Landmarks in this field
include Simon’s Administrative Behavior (1947), March and Simon’s Organizations (1958), and
Cyert and March’s Behavioral Theory of the Firm (1963), and Nelson and Winter’s An
Evolutionary Theory of Economic Change (1982). (Recent perspectives are given by March
1994 and Simon 1997.) This stream of research emphasizes that individuals typically have
limited powers of attention, memory, and calculation, and their decision-making behavior
therefore departs from the model of the perfectly rational expected-utility maximizer.
Boundedly rational agents satisfice rather than optimize—that is, they take the first available
alternative that meets some threshold of acceptability determined by heredity or experience.
They assume roles in organizations or social structures and then espouse values and employ
decision-making rules appropriate to those roles—but the same agent may play different roles,
depending on the decision context.9 They construct meaning and identity through their actions,
not merely obtain a desired set of consequences. Organizational decision processes are often
characterized by ambiguity rather than clarity of action, which is not always a bad thing—it may
be a source of innovation and adaptation to environmental change.
Although no one really disputes the premise that rationality is bounded, the membrane between
rational choice theory and boundedly-rational choice theory has remained only semi-permeable.
Rational choice theorists often add small amounts of noise to their models and/or use nonanalytic techniques such as simulation to investigate the behavior of less-than-perfectly-rational
agents, but they still mainly rely on parametric models of optimizing or quasi-optimizing
behavior by individual agents. Organization theorists occasionally use mathematical models of
9
The role-playing agent is sometimes referred to as homo sociologicus, in contrast to homo economicus. The latter
seeks to satisfy her own preferences, the former merely attends to her “station and its duties.” (Heap et al. 1992)
19
optimization, especially when describing the “technological core” of organizations, but otherwise
they tend to use a more interpretive vocabulary for describing organizational decision processes.
Behavioral decision theory and experimental economics. Behavioral experiments involving
preference and choice were performed by psychologists (e.g., Thurstone) as early as the 1930’s,
but they began to proliferate only after the publication of von Neumann and Morgenstern’s book,
as many researchers sought to test whether utility really was measurable and whether gametheoretic solution concepts really would predict the outcome of strategic contests. An excellent
review is given by Roth (1995); a popularized account of the early experiments is given by
Poundstone (1992). Behavioral decision theory has been used both to support and to challenge
the predictions of rational choice theory over the last 50 years. In laboratory problems involving
choices among simple lotteries, expected utility theory sometimes provides a parsimonious fit to
the data, but there are also predictable anomalies—e.g., reference point effects, distortions of
probabilities, attractions to sure things, etc.—that are better handled by other models such as
Tversky and Kahneman’s prospect theory. The existence of a “probability weighting function”
is now taken for granted by many behavioral researchers (Tversky and Fox 1995, Wu and
Gonzalez 1997, Prelec 1998). Meanwhile, in experimental games and markets, subjects
sometimes converge to rational equilibria and other times they do not. Simple models of
adaptive learning (e.g., Roth and Erev 1995, Camerer and Ho 1998) sometimes suffice to explain
the convergence when it occurs, and in other cases the market structure helps agents with low
intelligence to behave efficiently (e.g., Gode and Sunder 1993). But seemingly innocuous
institutional details (e.g., trading rules) can strongly affect the results, a fact which has been
exploited in the use of experimental markets to fine-tune the design of auction mechanisms.
Although behavioral research generally adopts a bounded-rationality perspective, it otherwise
hews to the standard of methodological individualism: the focus usually is on the behavior of
individuals under controlled conditions where they have little or no social interactions with
others. As such, behavioral experiments follow the scientific tradition of trying to strip a
phenomenon down to its barest essentials in order to measure things precisely and reveal the
operation of universal laws. But this strategy has proved somewhat less successful in revealing
laws of human behavior than laws of physics, or rather, its results are often more reminiscent of
quantum mechanics than classical mechanics. Human beings are accustomed to facing complex
problems, but their behavior does not necessarily become precise and regular when they are faced
with simpler ones. Decision processes, like other biologically-based phenomena, are contingent
on subtle features of the environment and the state of the organism. (See Einhorn and Hogarth
1981 and Payne et al. 1993 for discussions of contingent decision making. A broader view of the
role of contingency in natural science is given by Gould 1989.) When presented with a highly
stylized lottery or game problem, subjects do not attend only to the objectively given “rules.”
Their responses are determined in complicated ways by the entire context of the experiment: by
symmetries or asymmetries of the design, by the manner in which information is presented, by
the mode of response that is required, by limits on time and attention, by emotional reactions, by
social norms, by cooperative or skeptical attitudes toward the experimenter, and so on. The
uncertainty principle turns out, not surprisingly, to apply to psychological as well as physical
measurements: the very asking of a question and the language and context in which it is framed
trigger a constructive mental process that alters as well as reveals the state of the subject. This
20
constructive process often violates the “invariance principle,” namely that a choice between
alternatives should be invariant to the manner in which the alternatives are described, as long as
their objective attributes remain the same. Different descriptions often give rise to different
processes of mental construction, as partisan opinion pollsters are well aware. And when
experimental subjects are exposed to repeated trials of the same type of problem, they often learn
to behave in the manner they think the experimenter intends.
Austrian and subjectivist economics. The roots of rational choice theory extend back through
the marginalist revolution in economics in the late 19th Century, whose protagonists—Walras,
Jevons, and Menger—proposed a subjective theory of value based on ideas of utilitymaximization and competitive equilibrium. Menger’s work later became the cornerstone of a
distinctive Austrian school of economics (extended by Schumpeter, Hayek, von Mises, and most
recently Kirzner) that eventually parted company with neoclassical economics. To the Austrian
economists, the salient features of a free market economy are not the creation of surplus value
and the attainment of equilibria through routine production and exchange, but rather the
persistence of disequilibrating forces and the role of entrepreneurial “alertness” and “discovery”
in reacting to those forces. From their perspective, value is created when an alert individual
notices a discrepancy between prices in different markets or discovers an unexpected
technological innovation. Such opportunities for entrepreneurial profit are not merely
aberrations in an otherwise efficient market—they arise spontaneously and continuously and are
perceived as the very engine of economic progress and the wellspring of moral as well as
financial worth. The Austrian view of the market process is actually more radically
individualistic than the standard rational choice view. To the Austrians, the actors of rational
choice theory are mere computing machines who shuffle around their existing endowments of
resources to mutually increase their personal utility, but never create or discover anything new.
They only search for things that they already know are there to be found, and they find such
things with probabilities that they have estimated in advance. By comparison, the Austrians’
alert entreprenuer lives (and sometimes dies) by the exercise of her own free will. Her goal is to
find things whose prior existence is unknown and whose probability of discovery is therefore
unquantifiable. The fruits of her labor—the overlooked profit opportunities and new
technologies she is able to discover—are her own just reward, an ethic of “finders keepers.”
(Kirzner 1992) An even more radical position is held by Shackle, who emphasizes that the
enterprising individual does not merely discover the future, but invents it: “…we speak and
think as though decisions were creative. Our spontaneous and intuitive habits of mind treat a
decision as introducing a new strand into the tapestry as it is woven, as injecting into the dance
and play of events something essentially new, something not implicit in what had gone before.”
(Ford 1990, p. 290)
Stripped of its ideological baggage—and never mind whether the entrepreneur stands on anyone
else’s shoulders in order to see farther—the Austrian argument makes a valid point about the role
of “unknowledge” (what you don’t know that you don’t know) in the dynamics of economic and
social systems. The passage of time is experienced as a process of discovery, and decision
making, at its best, shapes the course of destiny. But in rational choice models, nothing ever
happens that the modeler has not foreseen, and there is indeed nothing for an intelligent agent to
21
do other than calculate how to best satisfy the desires she already possesses. The parameters of
the final equilibrium state are built into the initial conditions, and everyone knows it.
Evolutionary & complexity theory. Evolutionary arguments have long been used both to
attack and defend neoclassical economics and rational choice theory—an excellent survey is
given by Vromen (1995). The neoclassical evolutionary defense (Friedman 1953) merely asserts
that natural selection in a market economy will surely, somehow, in the long run, favor agents
who maximize over those who do not. (Of course, this argument assumes rather than proves that
maximizing behavior is both feasible and reproductively helpful.) A more sophisticated version
of the evolutionary defense, inspired by the work of Maynard Smith in evolutionary biology,
shifts the focus of the analysis from the individual to the population. Suppose that the population
consists of different types of boundedly rational agents, each type employing its own rules of
behavior governing interactions with other agents and with nature. If the “winner” of each
interaction is more likely to reproduce than the “loser,” then a mixed strategy equilibrium will be
attained in which the mixing probabilities refer to long-run frequencies of different types in the
population, rather than randomized choices by individuals. In this way, boundedly rational
agents might manage to behave rationally through demographics rather than cognition. A more
sophisticated version of the evolutionary defense shifts the focus of the analysis to “memes”
(units of behavior) that can be passed from one agent to another through instruction or imitation.
(Dawkins 1976, 1986, Dennett 1995). Memes that prove useful to the agents who carry them
will survive to be passed on to other agents. Eventually, most agents should end up equipped
with sets of memes that provide a good approximation to rational behavior—but meanwhile the
memes may have a life of their own, like Dawkins’ “selfish genes.”
Many researchers in the bounded rationality stream (e.g, Simon 1955, Nelson and Winter 1982)
have argued, to the contrary, that both biological and cultural evolution favor non-rational-choice
methods of reasoning, because the complexity of the environment overwhelms the optimizing
capabilities of both individuals and organizations. Support for the latter view also comes from
the emerging field of complexity theory identified with the Santa Fe Institute. Boundedly
rational agents who must adapt to complex, changing environments are better off employing
inductive reasoning rather than deductive reasoning (Arthur 1994), generating an array of
working hypotheses and weighting them on the basis of recent experience. As they acquire
expertise in their domain of specialization, they rely increasingly on pattern recognition and
automatic responses rather than conscious deliberation, and they find it hard to externalize their
thought processes or encode them in analytic models. They also band together with other agents
to form networks of “generative relationships” that lead to true innovation—i.e., the
implementation of strategies and creation of benefits not originally conceived (Lane et al. 1996).
The inductive decision rules and networks of relationships form rich “ecologies” that may exhibit
nonlinear, chaotic dynamics. Meanwhile, recent research in evolutionary psychology (Barkow,
Cosmides, and Tooby 1992) and cognitive neuroscience (Pinker 1997) suggests that the brain is
not a general-purpose reasoning machine that can be programmed with arbitrary cultural
software; rather, evolutionary forces operating on our Pleistocene ancestors may have produced
“hard-wired” modules in the brain for solving important types of perceptual and decision-making
tasks.
22
*
*
*
In the sequel, the threads of the alternative paradigms sketched above will be rewoven in a new
pattern. We will accept that human cognition is bounded in some ways and exquisitely subtle in
others, limiting the precision with which analytic models can describe or prescribe the behavior
of individuals; that social norms, roles, induction, and intuition can be useful determinants of
behavior; that individuals behave unimpressively when removed from their familiar surroundings
and isolated in laboratory cubicles; that they experience surprise and discovery with the passage
of time; and that they pool their cognitive resources in order to find creative solutions to complex
problems. At the same time, we will find qualified support for some of the key optimization and
equilibrium concepts of rational choice, even uncovering a deeper unity among the several
branches of the theory. But the philosophical and practical implications will turn out to be rather
different than those of the standard theory.
23
3.1 Arbitrage choice theory
This section presents the outline of arbitrage choice theory (ACT), a reformulation of rational
choice that circumvents the difficulties of the standard theory mentioned above. In some respects,
it avoids trouble merely by making weaker assumptions and not trying to predict exactly what
every agent will do in every situation. But in other respects it makes fundamentally different
assumptions about the nature of interactions among agents: what is knowable, how it comes to be
known, and what it means for agents to act rationally. The result is a theory that is much simpler
and more strongly unified than the standard one. It also leaves the door open for some
unorthodox speculation on the purpose of interactions between agents, which will be offered up
in a later section. The main assumptions are as follows:
Elements of arbitrage choice theory
1. Environment:
A set of agents
A set of events (states of nature and alternatives for agents)
Money (and perhaps other divisible, transferable commodities)
A contingent claims market
An outside observer
2. Behavior:
Physical behavior: choices among alternatives
Verbal behavior: small gambles or trades that agents are willing to accept
3. Rationality:
Additivity of acceptable gambles or trades
No ex post arbitrage
Environment: As before, the small-world environment includes one or more human agents and
a set of events that are under the control of nature and/or the agents: an event controlled by
nature is a state of nature and an event controlled by a human agent is an alternative.10 But there
are no “acts” in ACT—that is, there are no arbitrary, imaginary mappings of “consequences” to
events—there are only feasible alternatives. Indeed, the troublesome abstract consequences of
the standard theory do not appear here at all. Rather, we assume that the environment includes
money (a divisible, transferable, and universally desirable commodity) and perhaps other
commodities, and that there is a market in which contracts can be written to exchange money and
other commodities contingent on the outcomes of events. (The set of events is now essentially
defined by the possibility of writing contingent contracts on them.) Of course, agents’ choices
presumably do have consequences for them other than money and commodities, as in the
standard theory, but such consequences are not assumed to be directly observable or exactly
quantifiable: they are not explicitly modeled as part of the small world. Instead, they are
indirectly modeled through the effects that they exert on the agents’ willingness to trade
10
It is actually no longer necessary to distinguish a priori between events controlled by one player and events
controlled by another, nor, indeed, between alternatives and states of nature. The question of who controls what
need not be answered in order for the behavior of the players to be judged rational. What matters is the way in which
agents are willing to bet on the outcomes of events, which may reveal what they believe about who is in control. The
usual terminology nevertheless will be retained for convenience.
24
contingent claims to money and commodities. Finally, the environment includes an outside
observer, who merely embodies what the agents know about each other (and what we theorists
know about them). The observer’s role is one that can be played by any of the agents at any time.
Behavior: Two kinds of behavior are of interest, one of which is physical behavior—choices
among alternatives—as in the standard theory. The second kind of behavior will be loosely
called “verbal” behavior, and it determines the gambling or trading contracts that are written in
advance of the outcomes of events. In particular, verbal behavior consists of public offers by
agents to accept low-stakes contracts with specified terms, as well as counter-offers by other
agents to enforce (i.e., take the other side of) those contracts. Verbal behavior is never
counterfactual nor is it merely cheap talk: it has direct, though usually only incremental,
economic consequences for the agents. If one agent says “I’m willing to accept contract x,”
where x is a vector of amounts of money (and perhaps other commodities) to be exchanged as a
function of the outcomes of events, then the observer is free to choose a small non-negative
multiplier , and a contract will be enforced in which the vector of money and/or commodities
actually exchanged is x. (Both the offer x and the response  are considered as instances of
verbal behavior: communication is a two-way street.) For example, if the agent says “I would
give a dollar for a glass of lemonade if it’s hot today,” this means that she will pay $1 per glass
for any small (possibly fractional) quantity of lemonade at the discretion of anyone else, in the
event that the day is hot.11
The gambles or trades that agents are willing to accept play the same role that preferences play in
the standard theory, in the sense that they provide the rich set of ancillary measurements needed
for a quantitative theory of choice. However, the acceptance of a gamble or trade is not merely
synonymous with the existence of a preference. An agent who says “I would give a dollar for an
glass of lemonade” is not only saying that she prefers a glass of lemonade to a dollar in the usual
sense of the term. She is also saying she doesn’t mind who knows it, and furthermore she will still
trade the dollar for the lemonade even if someone else wants to take her up on it. Thus, verbal
behavior has an intersubjective quality that is not inherent in the concept of preference: it is a
process of reciprocal influence, not merely a process of thought.12 It has the potential to alter the
agents’ stakes in the outcomes of events, as small gambles may accrue into finite transfers of
wealth during preplay communication, and in this way it partly endogenizes the “rules of the
game” that the agents end up playing. (In the standard theory, by comparison, the rules of the
game are always fixed in advance and are not changed by the measurement of preferences.) By
construction, all verbal behavior is common knowledge in the sense of the specular common
knowledge of a public market, and no additional assumptions about common knowledge will be
needed later on.
It is also possible to allow statements about large and specific exchanges, such as “ I would give exactly ten
dollars for exactly one gallon of lemonade.” Statements of this kind are needed for modeling cooperative behavior,
enabling the agents to escape from prisoner’s dilemma games and other troublesome situations. But the scope of
this paper is limited to “noncooperative” behavior, for which small gambles are sufficient.
11
12
The type of verbal behavior described here was first introduced by de Finetti (1937) as a method of eliciting
subjective probabilities. De Finetti viewed its reliance on money and its game-theoretic overtones as potential
“shortcomings” vis-à-vis an axiomatization based on a qualitative binary relation. (footnote (a) in the 1964
translation). Here, these very qualities are viewed as its strengths vis-à-vis a binary preference relation.
25
A type of contract that is of special interest is a bet on the occurrence of one event conditional on
the occurrence of another event. An agent who says “There is at least a p chance of event E,
given event F” means that she will accept a contract in which she wins $1-p if E and F both
occur, loses $p if F occurs but not E, and wins or loses nothing if F does not occur, where p is a
number between 0 and 1. In other words, $p is the price she would be willing to pay for a lottery
ticket that returns $1 if E occurs and $0 otherwise, with the entire transaction to be called off if F
does not occur. Treating E and F as indicator variables as well as names for events, the payoff
for this contract can be expressed as $(E-p)F. The term “chance” henceforth will be used
exclusively to refer to a rate at which an agent is willing to bet money on the occurrence of an
event in this fashion. An agent is not expected to assign non-trivial chances to events under her
own control, namely her own choices. Thus, for example, we would not expect her to say “there
is a 50% chance I will choose alternative x.” (Or if she did, it would be foolish for anyone else to
bet with her.) However, the bets she accepts may legitimately depend on her own choices as
conditioning events. Thus, for example, she might say: “In the event that I choose alternative x,
the chance of event E is at least p.” There are several reasons why an agent might wish to
condition bets on her own choices. One is that the agent’s betting rate on an event might depend
in part on the value of a dollar to her if that event should occur. Insofar as her own choice may
affect how the value of a dollar depends on the outcomes of other events, her betting rate on an
event that is not under her control might reasonably depend on the outcome of an event that is
under her control. Another possibility is that the agent anticipates receiving information between
now and the time that she makes her choice, and the choice she ultimately makes therefore may
be a proxy for her future state of information. Insofar as betting rates may be legitimately
conditioned on the receipt of information, it is reasonable for an agent to condition a betting rate
on her own future choice. (We will see later that the trick of conditioning bets on one’s own
choices is essential to eliciting information about values as well as beliefs.) An agent’s
selection of conditioning events for her bets may, in this fashion, reveal what she believes to be
her set of alternatives—i.e., the set of events she believes she controls.
An agent could have a variety of motives for putting her money where her mouth is in the
fashion described above: she might wish to extract gains from trade, hedge her risks, express her
beliefs and values in a way that others will find credible, and/or influence the behavior of other
agents. Whatever the motivation, every such act of verbal behavior creates a small disturbance in
the environment as other agents hear the message and have the opportunity to respond. If another
agent responds affirmatively (e.g., by agreeing to sell a fraction of a lottery ticket or a glass of
contingent lemonade at the price named), this in turn perturbs the information state and financial
position of the first agent. She might decide to change the terms of subsequent contracts she
offers to accept, which in turn may trigger other responses (or not) on the part of other agents. If
the other agents all respond negatively (i.e., silently), this too may be informative to the first
agent: she might decide to make a bolder offer next time. The following scenario can be
envisioned: two or more agents encounter each other in a public place. At least one of them is
“active” in the sense that she proposes terms for gambling or trading contracts she is willing to
accept. At least one of the others plays the role of an observer who decides whether, and for
what stakes, to take the other side of such contracts. Perhaps the active agents initially proceed
with caution, offering only to “buy cheap” or “sell dear,” but after some testing of the waters they
26
begin to offer somewhat more attractive terms. Occasionally a deal is struck and a contract is
signed, slightly altering the financial positions and information states of the participants and
precipitating changes in the terms that the active agents offer for subsequent contracts. During
such a period of preplay communication, while messages and money are flying back and forth, it
is hard to be sure of any agent’s instantaneous state of information or intent. Our analysis
therefore will begin at a point when preplay communication has subsided, at least temporarily, so
that every active agent is putting forth her last, best offer but none is being taken. The agents are
eyeball to eyeball and no one is blinking. The observers (including ourselves) then proceed to
evaluate whether the active agents are behaving rationally.
It may appear that we have implicitly assumed the agents are expected-value or expected-utility
maximizers and that they have already reached a kind of competitive equilibrium in the
contingent claims market as the curtain rises on our analysis. Certainly that is one possible
interpretation that could be placed on the scenario just described. But in fact we have assumed
very much less. The detailed preference measurements and consistency assumptions that would
be needed to construct a full expected-utility representation have not been made, and the initial
and final endowments that would figure into an equilibrium analysis have not been observed. All
that is assumed known is that some of the agents—for whatever reason—are currently offering to
accept various gambling or trading contracts, and the rest of the agents—for whatever reason—
are not responding to those offers. A kind of equilibrium exists, but it is mainly an equilibrium
of knowledge: it is common knowledge, in a practical sense, that various gambles or trades are
being offered and that there are no takers.
Rationality: The most striking difference between the present theory and the standard one is the
radical streamlining of the assumptions about rationality. There are only two such assumptions:
additivity and no ex post arbitrage. The additivity assumption is more structural than
substantive: it means that an active agent’s willingness to accept a contract is unaffected by an
observer’s simultaneous enforcement (or not) of other contracts with the same agent or other
agents. This assumption is justified in part by the smallness of the stakes and in part by the fact
that the preplay communication process is assumed to have converged. An observer’s
enforcement of any one contract at this point does not appreciably alter the distribution of wealth
or information, and hence it should not dull anyone’s appetite for other contracts they are
simultaneously offering to accept. But it is also partly justified as being merely a rule of the
game: the agents are aware that the contracts they offer to accept may be bundled together by an
observer, so they should only accept contracts that are desirable under those conditions. No
assumption is made concerning the completeness of verbal behavior—i.e., the number or scope
of the contracts that agents must accept. They need not offer prices or exchange rates for any
commodities or contingent claims, nor are they required to name prices at which they would
indifferently buy or sell. Since they are not required to accept any contracts at all, they are not
unduly burdened by the requirement that acceptable contracts should be additive.
The substantive rationality postulate is the no-arbitrage axiom, which embodies rationality as
judged from the perspective of an observer. Recall that the observer’s role is one that can be
played by any of the active agents: he knows only what is common knowledge and he trades
only what is commonly traded. We assume the observer to be completely naïve concerning the
27
likelihoods of events, and consequently he is interested only in opportunities for riskless profit—
that is, arbitrage. An arbitrage opportunity is a collection of acceptable contracts that wins
money for the observer (and loses money for the agents in the aggregate) in at least one possible
outcome of events, with no risk of loss for the observer in any outcome. There is ex post
arbitrage if the outcome that occurs is one in which the observer could have won a positive
amount of money without placing any money at risk. The principal axiom of rationality, then, is
that there should be no ex post arbitrage. In this way, rationality depends on both verbal and
physical behavior—whether what the players do is sufficiently consistent with what they say in
order to avoid material exploitation by the naïve observer—and it is evaluated posterior to the
outcomes of events. This is the only consistency condition imposed by the theory, and it refers
only to “primal” variables—i.e., to observable behavior. By comparison, the standard theory
imposes many different kinds of consistency conditions, some of which refer to primal variables
such as choices and preferences and others of which refer to dual variables such as probabilities
and utilities.
The ex post evaluation of rationality is essential for dealing with situations in which the agents
assign chances of 0 or 1 to events. If the players jointly accept contracts that are equivalent to
assigning a 100% chance to an event E that (they believe) is controlled by nature, then they are
deemed irrational if event E fails to occur, so they should accept such contracts only if they are
sure that E will occur. If they jointly accept contracts that are equivalent to assigning a 100%
chance to an event E that (they believe) is under their own control, then they had better make sure
that E occurs. Whether the events are really under their control is irrelevant to the observer. The
players are deemed irrational after the fact if they should have known better or acted otherwise.
3.2 Fundamental theorem and examples
On the surface, the preceding assumptions about environment, behavior, and rationality merely
appear to describe the operation of a fairly normal contingent claims market or oddsmaking
operation. As such, it would not be surprising to find that we can reproduce some basic results of
subjective probability and financial economics—e.g., theorems on coherent odds and asset
pricing by arbitrage. There is also a close connection between the no-arbitrage standard of
rationality and the familiar concept of Pareto efficiency: as noted earlier, a commonly known
Pareto inefficiency is just an arbitrage opportunity under another name. Hence we might also
expect to derive some basic results of welfare economics. What is perhaps not so apparent is that
the assumptions given above are also sufficient to replace much of the remaining apparatus of
standard rational choice theory—the preference axioms, knowledge assumptions, and
equilibrium concepts—and to permit the analysis not only of markets but also games and
personal decisions. This is accomplished by invoking a single mathematical theorem, namely a
familiar duality theorem (a.k.a. separating-hyperplane theorem, or “theorem of the alternative”)
of linear algebra.
THEOREM 0: Let X be an mn matrix, let  be an m-vector, let  be an n-vector, and let
[X](s) and (s) denote the sth elements of X and , respectively. Then exactly one of the
following systems of inequalities has a solution:
28
(i)
  0, X  0, [X](s) < 0
(ii)
  0, X  0, (s) > 0
In applications, the columns of the matrix X will be identified with different outcomes or
commodities. (More generally, a column might correspond to a state-contingent commodity.)
The rows of X will be identified with gambles or trades that agents have offered to accept. If xi
denotes the ith row of X, then xi(s) is the monetary payoff in state s yielded by the ith gamble, or
the quantity of commodity s yielded by the ith trade. System (i) is the primal characterization of
irrational behavior in terms of arbitrage: if (i) has a solution, then there is a weighted sum of the
gambles or trades in X, with weight vector , leading to arbitrage in state s (or in commodity s) .
System (ii) is the dual characterization of rational behavior in terms of probabilities, prices,
and/or equilibrium conditions: if (ii) has a solution, then there is a vector  of probabilities or
prices that rationalizes all the gambles or trades that have been accepted, in the sense that it
assigns them all non-negative expected value or profit, and in which state s has strictly positive
probability or commodity s has a strictly positive price. By simply varying the structure of the X
matrix, the preceding theorem becomes, successively, the fundamental theorem of welfare
economics, the fundamental theorem of subjective probability, the fundamental theorem of
decision analysis, the fundamental theorem of noncooperative games, the fundamental theorem
of asset pricing, and more. The fact that all these results are variations on the same no-arbitrage
theme, and the same duality theorem, is illustrated by the following simple examples. (For the
additional technical details, see Nau and McCardle 1990 & 1991; Nau 1992b, 1995abc.)
Example 1: pure exchange
The most interesting applications of arbitrage choice theory are to situations involving
uncertainty. Yet the simplest illustration is given by a problem of pure exchange, such as the
following:
Alice: I will trade 4 apples for every 3 bananas, or 2 bananas for every 3 apples.
Bob: I will trade 2 apples for every 1 banana, or one banana for every 3 apples.
If fruit serves as money in this primitive economy, then Alice and Bob are irrational. For, Bob
will give us 4 apples in exchange for 2 bananas, and we can then give 3 of the apples to Alice
and get our 2 bananas back, for an arbitrage profit of one apple. The situation can be illustrated
by a familiar Edgeworth box diagram:
<= Bob's bananas
Alice's bananas =>
<= Bob's apples
AB
B'
A'
29
Alice's apples =>
The agents’ current endowments are represented by the dot labeled AB, and the lines passing
through it (solid for Alice and dashed for Bob) represent neighboring endowments to which they
would be willing to move by trading apples for bananas or vice versa. Assuming that more of
any fruit is preferred to less, Alice prefers any endowment on or above the solid curve to her
current endowment, and Bob prefers any endowment on or below the dashed curve to his current
endowment. These are analogous to “indifference curves” in standard welfare analysis, although
here they need not represent strict indifference—let us call them “acceptable-trade curves”
instead. An arbitrage opportunity exists because, for example, Alice is willing to move to the
point A, and Bob is willing to move to the point B, at which their total number of bananas is
conserved but they have one fewer apple between them.
The necessary and sufficient condition for avoiding arbitrage in this situation is that the
acceptable-trade curves passing through the endowment point should not cross: they should be
separated by a straight line passing through that point. (If the curves were smooth rather than
kinked, they would have to be tangent.) A similar, more general result applies to any number of
agents and any number of commodities, although with more than two agents and two
commodities we can no longer draw an Edgeworth box. Instead, we may plot the cone generated
by the vectors of acceptable commodity trades, as shown in the diagram below. Alice’s two
acceptable trades are the vectors labeled A1 and A2 (trade A1 is “minus 3 apples, plus 2 bananas,”
etc.), and Bob’s two acceptable trades are labeled B1 and B2. (These vectors are shown as rays
emanating from the origin.) Acceptable trades may non-negatively combined, and the question is
whether there is any non-negative linear combination that yields a strictly negative vector, which
is an arbitrage opportunity. In this case, A2+(5/3)B1 is strictly negative, yielding -1/3 apple and
-1/3 banana to the two agents as a group.
A1
Bananas
B1
Apples
A2+(5/3)B1
B2
A2
The general result is that there is no arbitrage opportunity if and only if there is a hyperplane
passing through the origin that separates all the acceptable trade vectors from the negative orthant
30
(the lower left quadrant in the two-dimensional case shown here). The normal vector of such a
hyperplane defines relative prices for the commodities under which the net value of every
acceptable trade is non-negative. Under such a price vector it is as if every agent’s endowment is
already “optimal” for her under her “budget constraint”: she cannot make herself better off than
she already is by buying and selling commodities at those prices. More precisely, it is impossible
for her to cash in part of her current endowment and use the proceeds to make a purchase that
replicates a trade she is willing to accept, with some money left over. In this sense, the current
endowments constitute a competitive equilibrium with respect to the prices determined by the
separating hyperplane. We therefore have the following result:
THEOREM 1: Agents who trade publicly in an exchange economy are rational (avoid
arbitrage) if and only if their (final) endowments are a competitive equilibrium with
respect to a system of non-negative commodity prices.
Proof: Let X be the matrix whose columns are indexed by commodities and whose rows
correspond to acceptable trades, where the sth element of row i, xi(s), is the quantity of
commodity s received by an agent in the ith trade.13 By Theorem 0, either there is a
weighted sum of trades that yields an arbitrage profit (i.e., a negative aggregate amount of
money and non-positive aggregate amounts of all other commodities to the agents as a
group), or else there is a vector of non-negative relative prices, with money as the
numeraire, under which every trade has non-negative value for the agent.
This is of course just a combination of the first and second theorems of welfare economics with
the no-arbitrage condition taking the place of the familiar Pareto optimality condition, to which it
becomes equivalent when the concept of preference is is replaced by public willingness to trade.
(A more detailed version of this result is given in Nau and McCardle 1991.)
This example, though elementary, illustrates some generic features of arbitrage models. First,
the informational demands are relatively low. We do not observe the agents’ actual endowments:
they might be long or short in the fruit market, and we have no way to audit their books. Nor do
we presume to know how their trading behavior would change if their endowments had some
other, hypothetical values. All we know are the trades they are willing to make in the vicinity of
their current endowment, whatever that might be. Of course, with lower informational demands
our predictive ambitions must also be lower. Following Edgeworth, we do not try to predict the
transition from an initial disequilibrium state to a final equilibrium state, which is intrinsically
indeterminate. We merely conclude that it is rational for the agents to end up in an equilibrium
state, regardless of where they may have started.
13
In the example, the matrix X looks like this:
Apples
Trade #1 for Alice (A1)
-4
Trade #2 for Alice (A2)
3
Trade #1 for Bob (B1)
-2
Trade #2 for Bob (B2)
3
Bananas
3
-2
1
-1
31
Second, the acceptable-trade curves are generally kinked at the current endowment. Thus, agents
do not necessarily trade in both directions at the same rates: they maintain some kind of bid-ask
spread. This feature is characteristic of most real markets: it gives the agents the ability to build
in a profit margin—an incentive to engage in trade in the first place—and to hedge the risks of
exploitation by others who may have better information. It very importantly changes the
interpretation of the trading or betting rates: they are not merely rates at which the agent is
indifferent to trading, they are rates that the agent doesn’t mind revealing and which she will
honor even upon discovering that someone else is willing to take the other side.
Third, the labeling of the agents is not essential to the analysis. We don’t really care which agent
accepts which trade: all that matters is the set of trades that are known to be acceptable to
someone. For example, it is unimportant that the agent who accepts trade A1 is the “same” agent
who also accepts trade A2. This means we do not need to assume a complete preference ordering
for each agent before the analysis can get off the ground. We don’t necessarily care what the
agents may think privately: we are only concerned with what happens when they interact with
other agents (or ourselves) in a public place.
Last but not least, rationality or irrationality is inherently a property of the group. Notice that
Alice by herself is perfectly rational, as is Bob, but as a group they are irrational. However, the
irrationality of the group rubs off on all its members. Since their acceptable trades are common
knowledge by construction, they should notice that their joint behavior creates an arbitrage
opportunity for someone, and they would be foolish not to exploit it for themselves. Thus, it is
irrational for Alice to ask for only three apples in exchange for two bananas when Bob is already
offering four apples for two bananas, and similarly for Bob. Presumably upon noticing the
inconsistency one or both will revise their trading rates, perhaps after some actual trades have
taken place.
Example 2: elicitation and aggregation of belief
The next example illustrates how the basic model extends from conditions of certainty to
uncertainty when the trades are contingent on events. Suppose that a geological test is to be
conducted prior to drilling an oil well, and three participants make the following statements:
Alice: The chance of striking oil here is at least 40%.
Bob: If the geological test turns out to be negative, the chance of striking oil can’t be any
more than 25%.
Carol: Even if there’s oil here, the chance of a negative test result is at least 50%.
Remember that the term “chance” is used throughout to refer to rates at which the agents are
willing to bet money on the conditional or unconditional outcomes of events. Here, for example,
Alice indicates that she is willing to bet on the event that oil is found at odds of 40:60 in favor.
In other words, she will accept a small bet with net payoffs in the following proportions:
32
Payoff to Alice
Positive, Oil
$3
Positive, No Oil
-$2
Negative, Oil
$3
Negative, No Oil
-$2
(Note that this table does not have the same interpretation as the tables of acts and alternatives in
Tables 1-3. The entries in the cells are neither utilities nor final consequences: they are merely
small quantities of money received in addition to whatever other consequences are attached to
events.) Meanwhile, Bob indicates that he is willing to bet against finding oil at odds of 75:25
against, given a negative test result. In other words, Bob will accept a small bet with net payoffs
in the proportions:
Payoff to Bob
Positive, Oil
$0
Positive, No Oil
$0
Negative, Oil
-$3
Negative, No Oil
$1
(Note that Bob’s bet against finding oil is called off—i.e., the net payoff is zero—if the test result
is positive.) Finally, Carol indicates that she is willing to bet on a negative test result at 50:50
odds, given that oil is found, meaning she will accept a small bet with net payoffs proportional
to:
Payoff to Carol
Positive, Oil
-$1
Positive, No Oil
$0
Negative, Oil
$1
Negative, No Oil
$0
This dialog provides an example of verbal behavior that conveys information about beliefs.
Alice (who is perhaps a geologist familiar with the area) has given information about her prior
belief concerning the presence of oil. Bob (who is perhaps an oil driller with some experience in
using the geological test under similar conditions) has given information about his posterior
belief in the presence of oil, given a negative test result. Carol (who is perhaps a technician
familiar with properties of the test) has given information about her belief concerning a “false
negative” test result. Note that no single agent has given a very complete account of his or her
beliefs: each has given a bound on a single conditional or unconditional betting rate. No one
claims to know everything about the situation, and we are not in a position to judge whether any
individual’s beliefs are coherent. Yet, assuming that the geological test is going to be conducted
and that drilling will occur at least if the result is positive, we can still reach the following
conclusion:
Alice, Bob, and Carol are irrational in the event that the test result is positive and no oil
is found upon drilling.
To see this, note that an observer can make bets with the agents in which the payoffs to the
players under the various possible outcomes of events are as follows:
Payoff to Alice
Payoff to Bob (x2)
Payoff to Carol (x3)
Total
Positive, Oil
$3
$0
-$3
$0
Positive, No Oil
-$2
$0
$0
-$2
33
Negative, Oil
$3
-$6
$3
$0
Negative, No Oil
-$2
$2
$0
$0
(The observer has taken the liberty of applying appropriate multipliers to the bets with Bob and
Carol: the agents offered to accept any small bets with net payoffs proportional to those shown
earlier, with non-negative multipliers to be chosen at the discretion of the observer.) A negative
aggregate payoff for the agents is a profit for the observer. The observer is seen to earn a riskless
profit if no oil is found following a positive test result, 14 in which case he would be justified in
considering the players to have behaved irrationally.
The usual objection raised at this point is that even if the observer earns a riskless profit, it does
necessarily mean that any individual has behaved irrationally. Each will merely blame the others
for betting at the wrong odds, thereby creating the observer’s windfall. The players are entitled
to their subjective beliefs, and their own bets may be rational by their own lights. But the
rejoinder to this objection is that they are all guilty of irrationality because (as in the previous
example) they are all witness to all of the bets: any one of them can exploit the same opportunity
as the observer. For example, suppose Alice takes the opposite side of the bets with Bob and
Carol in the table above. Her payoff then would be as follows:
Payoff from Bob (x2)
Payoff from Carol (x3)
Total
Positive,
Oil
$0
$3
$3
Positive,
No Oil
$0
$0
$0
Negative,
Oil
$6
-$3
$3
Negative,
No Oil
-$2
$0
-$2
The bottom line is the same bet she has already offered to accept—except that it yields $2 more if
no oil is found after a positive test result! Meanwhile, Bob and Carol can do similarly. So it is
as if the players all see a piece of paper lying on the ground saying “Get the bet of your choice
and receive an extra $2 if no oil is found following a positive test result.” If they all decline to
pick it up, they evidently all believe it won’t pay off—and if it doesn’t pay off, that apparent
belief is vindicated. But if it does pay off, they are all irrational ex post. Note that if any agent
had attempted to exploit the arbitrage opportunity, the opportunity presumably would have
evaporated: she would have raised the stakes until a point was reached at which she or some of
the other agents would have wished to stop betting at the same rates. In so doing, she would
have changed the rules of the game.
Of course, there is a theorem which spells out the conditions under which the players are rational.
It is the fundamental theorem of subjective probability (de Finetti 1937, 1974), otherwise known
as the Dutch Book Theorem:
14
In order to instantiate all of the gambles, it is necessary to know the outcomes of both events: performing the test
and drilling for oil. However, if the test is performed but no drilling is performed following a negative result, then
all of the gambles must be called off, and the aggregate payoff is the same as if drilling had occurred, namely zero.
Hence, from the observer’s perspective, it suffices that the test will be conducted and that drilling will occur if the
result is positive.
34
THEOREM 2: Bettors are rational (avoid ex post arbitrage) if and only if there is a
probability distribution on outcomes that assigns non-negative expected value to every
acceptable bet and assigns positive probability to the outcome that occurs.
Proof: Let X be the matrix whose rows are payoff vectors of gambles the players have
agreed to accept. Then by Theorem 0, either there is ex post arbitrage against outcome s
or else there is a probability distribution assigning positive probability to outcome s and
assigning non-negative expectation to all acceptable bets.
In the example, ex post arbitrage is possible against the outcome {Positive, No Oil}. According
to the theorem, every probability distribution on outcomes that assigns non-negative expectation
to all the players’ bets must must assign zero probability that outcome. In fact, there is a unique
probability distribution that rationalizes the bets in this fashion, namely:
Probability
Positive, Oil
20%
Positive, No Oil
0
Negative, Oil
20%
Negative, No Oil
60%
From the observer’s perspective, therefore, it is as if the players have revealed a unique
subjective probability distribution on outcomes. The observer doesn’t care who has accepted
which bet: everyone’s money is equally good. So, the bets might just as well have been
accepted by a single “representative agent” with a unique probability distribution assigning zero
probability to the event {Positive, No Oil}.
The probability distribution of the representative agent has the appearance of an aggregate
probability distribution for the agents, but caution must be exercised in interpreting it that way.
First, the rates at which agents are willing to bet money are influenced not only by their
judgments of the relative likelihood of events but also by their relative appreciation of money in
different events. Other things being equal, an agent would rather bet money on an event in
which, were it to occur, money would be worth more than in other events. Indeed, if we suppose
that the agents are subjective expected utility maximizers, then their betting rates ought to equal
the renormalized product of their true probabilities and relative marginal utilities for money—
quantities that are called “risk neutral probabilities” in the finance literature.15 We have not
axiomatized SEU at this point, so the interpretation of betting rates in terms of subjective
probabilities and marginal utilities for money is merely suggestive. However, the fact that
betting rates may not be “true” probabilities is not necessarily problematic: betting rates are
actually more useful than probability judgments because they suffice to determine monetary
gambles that the agents will accept.
15
Bob might become wealthy if he drills and strikes oil and nearly bankrupt if he drills and does not strike oil.
Money earned from bets therefore might have higher marginal utility for him if he does not strike oil. For example,
suppose that Bob’s true posterior probability of striking oil after a negative test result is 50%, but his marginal utility
for money is three times higher in the event of no oil. Then his risk neutral probability for not striking oil will be
three times as great as his risk neutral probability for oil.
35
Another caveat in interpreting the representative agent’s probability distribution is that individual
agents are not required to assert sharp bounds on betting rates: they are not required to post
betting rates on all events, nor are they required to post rates at which they will bet indifferently
in either direction. Hence, the betting rates of any single agent need not determine a unique
supporting probability distribution, even if the representative agent’s distribution is unique (as it
is in this case). In general, the betting rates offered by a single rational agent will be consistent
with some convex set of probability distributions, and the intersection of all such sets is the set of
probability distributions characterizing the representative agent (Nau and McCardle 1991).
The situation described in this example resembles a competitive equilibrium in a contingent
claims market, in which the probability distribution of the representative agent is a vector of
competitive prices for Arrow-Debreu securities in the absence of discounting. (See, for example,
Drèze 1970.) As such, it could easily be embellished with additional state-contingent
commodities that the agents might trade among themselves (e.g., oil futures), in which case noarbitrage would imply the existence of a (not-necessarily-unique) vector of supporting prices for
all commodities in all states. Again, however, it should be kept in mind that the usual
prerequisites of an equilibrium analysis have not been assumed: we do not have a detailed model
of the beliefs, preferences, or prior endowments of any agent.
This example has illustrated some basic properties of arbitrage models for problems involving
uncertainty. The behavior of the agents is consistent with expected utility maximization and the
existence of an equilibrium, but many of the usual details are lacking. For one thing, we do not
know anyone’s “true” subjective probabilities. Instead, we observe quantities that behave like
probabilities but must be interpreted as amalgams of true probabilities and marginal utilities for
money—quantities that are known as “risk-neutral probabilities.” Furthermore, the risk neutral
probability distribution of any single agent need not be uniquely determined: there may be a
convex set of such distributions, each of which is consistent with the bets the agent has offered to
accept in the sense of assigning them all non-negative expected value. Finally, the risk neutral
probability distribution of the group—i.e., of the representative agent—is generally more
precisely determined than those of the individual agents. In this sense the group is “more
rational” than any of its constituents.
Example 3: decisions under uncertainty
The preceding example was a model of (almost) pure belief: the only events were states of
nature. The next example illustrates a decision problem under uncertainty: a situation in which
some events are states of nature, all other events are alternatives for a single agent, and the
problem is for that agent to rationally choose among those alternatives. Suppose that the event
under the control of an agent is whether she carries an umbrella, the event not under her control
is whether it rains, and she makes the following statement:
Alice: “I’d carry my umbrella today only if I thought the chance of rain was at least 50%,
and I’d leave it at home only if I thought the chance was no more than 50%.”
In other words, Alice is willing to accept either or both of the following bets:
36
Table 6:
Payoff to Alice for bet #1
Payoff to Alice for bet #2
Umbrella,
Rain
$1
$0
Umbrella,
No rain
-$1
$0
No umbrella,
Rain
$0
-$1
No umbrella,
No rain
$0
$1
Note that these bets depend on Alice’s choice (only) as a conditioning event: given that she ends
up carrying her umbrella, she will bet on the occurrence of rain at 50:50 odds, and given that she
ends up not carrying her umbrella, she will bet against the occurrence of rain at 50:50 odds. Bets
of this kind do not reveal any information about Alice’s (apparent) probabilities for states of
nature. Rather, they reveal information about Alice’s (apparent) utilities for outcomes of
events.16 Assuming money is equally valuable in all outcomes, it is as if her utility function for
outcomes has the following form:
Table 7a
Umbrella
No Umbrella
Rain
1
-1
No Rain
-1
1
To see the correspondence between the acceptable bets in Table 6 and the utility function in
Table 7a, note that if Alice has constant marginal utility for money, the acceptable bets imply that
she will carry her umbrella or not according to whether she thinks the probability of rain is
greater or less than 50%, which is exactly the same behavior that is implied by the utility
function. Table 7a does have the same interpretation as Table 3: the entries in the cells are
utility values. However, as with any utility functions, the values are not uniquely determined:
the origin and scale are arbitrary, and an arbitrary constant may also be added to each column
without affecting comparisons of expected utility between alternatives. Thus, for example, an
equivalent utility function that better represents the consequences described in section 2.1 is:
Table 7b
Umbrella
No Umbrella
Rain
0
-2
No Rain
1
3
Of course we do not assume, a priori, that Alice is an expected-utility maximizer. She may have
any reasons whatever for accepting the bets summarized in Table 6. The implied utility functions
of Tables 7ab are merely suggestive interpretations of her behavior. But it is worth noting that if
Alice does think in the fashion of expected-utility analysis, and if she doesn’t mind who knows
her utilities, then it is in her own interest to accept the bets of Table 6. Regardless of the
probability that she may assign to the event of rain prior to making her choice, the bets in Table 6
cannot decrease her total expected utility. Indeed, unless she ends up perfectly indifferent
The revelation of utilities via bets that are conditioned on the agents’ own choices is discussed by Nau and
McCardle (1990) and Nau (1992b, 1995c). A systematic approach to the elicitation of such bets is described by Nau
(1995b).
16
37
between carrying the umbrella or not, they can only strictly increase her total expected utility.
The defining quality of the bets in Table 6, under an expected-utility interpretation, is that they
amplify whatever differences in expected utility Alice perceives between her two alternatives.
For example, if she later concludes that carrying the umbrella has higher expected utility than not
carrying it because her probability of rain turns out to be greater than 50%, then she will carry the
umbrella and bet #1 will be in force—and bet #1 yields positive marginal utility precisely in the
case that her probability of rain is greater than 50%.
Now suppose that at the same time and place that Alice is making her statement, a second agent
is saying the following:
Bob: “I think the chance of rain is at least 75%—whether or not you carry your
umbrella!”
In other words, Bob is willing to accept either or both of the following bets:
Payoff to Bob for bet #1
Payoff to Bob for bet #2
Umbrella,
Rain
$1
$0
Umbrella,
No rain
-$3
$0
No umbrella,
Rain
$0
$1
No umbrella,
No rain
$0
-$3
This statement reveals information about Bob’s beliefs: assuming that money is equally valuable
to him whether or not it rains, his probability of rain is evidently at least 75%, and furthermore he
does not regard Alice’s behavior as informative. (Perhaps Bob is the local weatherman.)
We are now in a position to predict (or perhaps prescribe) what Alice should do, namely: she
should carry the umbrella, because otherwise there is ex post arbitrage. If an observer takes bet
#2 with Alice (scaling it up by a factor of two) and bet #2 with Bob, the result is as follows:
Payoff to Alice for bet #2 (x2)
Payoff to Bob for bet #2
Total
Umbrella,
Rain
$0
$0
$0
Umbrella,
No rain
$0
$0
$0
No umbrella,
Rain
-$2
$1
-$1
No umbrella,
No rain
$2
-$3
-$1
Hence, the observer earns a riskless profit in the event that no umbrella is carried.
As in the previous example, the question might be asked: why should Alice be bound by Bob’s
beliefs concerning the likelihood of rain? Perhaps she believes the probability of rain is less than
50% even after hearing Bob’s statement, in which case, based on the utility function inferred
from her own previous testimony, she should not carry the umbrella. But if this is so, she ought
to hedge her risks by betting with Bob, and she should keep betting with him and raising the
stakes until someone adjusts his or her betting rate to eliminate the arbitrage opportunity in the
“no umbrella” event. (And the same for Bob with respect to Alice.) If, instead, each agent hears
the other’s statement but does not respond to it, they evidently are both certain that Alice will
38
carry the umbrella. Of course, to an observer, it doesn’t matter whether one agent or two is
speaking. Bob’s statement can just as well be made by Alice in the instant before she decides
whether to carry the umbrella, in which case the conclusion is the same: she should take the
umbrella.17
Note that decision analysis was carried out in a novel fashion in this example: rather than asking
the agent to articulate a probability distribution and utility function and then advising her to
choose the alternative with the highest expected utility, we asked the agent—and those around
her!—to articulate the gambles they were willing to accept and then advised her to choose the
alternative that avoided ex post arbitrage. Of course, there is a theorem that says the two
approaches are equivalent, with one important exception: the latter method does not require the
explicit separation of probability from utility, and it does not assume that the agent’s utilities are
state-independent (Nau 1995b).
THEOREM 3: A decision maker is rational (avoids ex post arbitrage) if and only if
there exists a probability distribution on states of nature and a (not-necessarily-stateindependent) utility function such that every gamble she accepts yields a non-negative
increment of expected utility, the alternative she chooses maximizes expected utility, and
the state that occurs has positive probability.
Proof: This result is a hybrid of Theorems 2 (above) and 4 (below). More details can be
found in Nau (1995b).
Example 4: games of strategy
The next example illustrates the generalization from a single-agent game against nature to a
multiple-agent game of strategy. Contrary to von Neumann and Morgenstern, it entails no
additional modeling assumptions or rationality concepts.18
Alice: “I’d carry my umbrella only if I thought there was at least a 50% chance that Bob
would dump a bucket of water out the window as I walked by, and I wouldn’t carry it
only if I thought the chance was less than 50%.”
Interestingly, if Bob’s statement about the chance of rain is made unconditionally—i.e., if he merely says “I think
the chance of rain is at least 75%”—then there is no arbitrage opportunity. Any bet with Bob then has a positive
payoff for him in the outcome {Umbrella, Rain}, but no bet with Alice has a negative payoff for her in the same
outcome that could be used to hedge the observer’s risk. In such a scenario, it is possible that Alice’s eventual
decision as to whether to carry the umbrella will be based on the receipt of private information previously unknown
to Bob, and Bob is tacitly admitting this possibility by failing to condition his statement on Alice’s behavior.
17
Aumann (1987) also observes that is is possible to “[do] away with the dichotomy usually perceived between the
‘Bayesian’ and the ‘game-theoretic’ view of the world.” Our demonstration of this fact is, roughly speaking, the dual
of Aumann’s, but it does not require the assumption of information partitions on a larger set of events nor the
assumption of a common prior distribution, and it also does away with the other perceived dichotomy between
strategic and competitive behavior.
18
39
Bob is now cast in the role of rainmaker rather than weatherman, and Alice’s implied utility
function is the same as before (Tables 7ab), with the event label “Rain” merely replaced by
“Dump.” But suppose that Bob (unlike nature) has a malicious interest in getting Alice wet, as
indicated by the following claim:
Bob: “I would dump a bucket of water out the window as Alice walked by only if I
thought there was at least a 50% chance she wasn’t carrying her umbrella, and I wouldn’t
dump it only if I thought otherwise.”
In other words, Bob will accept the following bets:
Payoff to Bob for bet #1
Payoff to Bob for bet #2
Umbrella,
Dump
-$1
$0
Umbrella,
No dump
$0
$1
No umbrella,
Dump
$1
$0
No umbrella,
No dump
$0
-$1
Bob is now revealing information about his relative utilities for outcomes, not his beliefs. It is as
if his utility function is of the form:
Umbrella
No umbrella
Dump
-1
1
No dump
1
-1
…because (only) someone with a utility function equivalent to this one would accept the bets that
Bob has accepted (assuming constant marginal utility for money). Putting Alice’s and Bob’s
apparent utility functions together, we find it is as if they are players in a noncooperative game
with the payoff matrix:
Umbrella
No umbrella
Dump
1, -1
-1, 1
No dump
-1, 1
1, -1
where the numbers in the cells are the utilities for Alice and Bob respectively. This game was
introduced earlier (Table 4) as “matching pennies,” and it has a unique Nash equilibrium in
which both players randomly choose among their two alternatives with equal probabilities.
What prediction of the outcome of the game can be made by arbitrage arguments? The set of all
acceptable gambles is now as follows:
Payoff to Alice for bet #1
Payoff to Alice for bet #2
Payoff to Bob for bet #1
Umbrella,
Dump
$1
$0
-$1
Umbrella,
No dump
-$1
$0
$0
40
No umbrella,
Dump
$0
-$1
$1
No umbrella,
No dump
$0
$1
$0
Payoff to Bob for bet #2
$0
$1
$0
-$1
As it happens, ex post arbitrage is not possible in any outcome, so Alice and Bob may do
whatever they please. However, from an observer’s perspective, Alice and Bob appear to believe
that they are implementing the Nash equilibrium solution—that is, they appear to believe that all
four outcomes are equally likely! To see this, note that the observer can rescale and add up the
gambles in the following way:
Payoff to Alice for bet #1 (x4)
Payoff to Alice for bet #2 (x2)
Payoff to Bob for bet #1 (x1)
Payoff to Bob for bet #2 (x3)
Total
Umbrella,
Dump
$4
$0
-$1
$0
$3
Umbrella,
No dump
-$4
$0
$0
$3
-$1
No umbrella,
Dump
$0
-$2
$1
$0
-$1
No umbrella,
No dump
$0
$2
$0
-$3
-$1
The bottom line is equivalent to betting on the outcome {Umbrella, Dump} at odds of 1:3 in
favor—i.e., betting as if the probability of this outcome is at least 25%. Alternatively, the
gambles can be combined this way:
Payoff to Alice for bet #1 (x0)
Payoff to Alice for bet #2 (x2)
Payoff to Bob for bet #1 (x3)
Payoff to Bob for bet #2 (x1)
Total
Umbrella,
Dump
$0
$0
-$3
$0
-$3
Umbrella,
No dump
$0
$0
$0
$1
$1
No umbrella,
Dump
$0
-$2
$3
$0
$1
No umbrella,
No dump
$0
$2
$0
-$1
$1
which is equivalent to betting as if the probability of {Umbrella, Dump} is no more than 25%.
Hence, between them, Alice and Bob appear to believe the probability of {Umbrella, Dump} is
exactly 25%—and of course by symmetry the same trick can be played with all the other
outcomes.
What is remarkable about this example is that neither Alice nor Bob has revealed any
information whatever about his or her beliefs—they have merely revealed information about their
values via appropriate gambles. Yet this turns out to be operationally equivalent to asserting
beliefs that correspond to a Nash equilibrium! Why did this happen? Of course there is another
theorem lurking around, and it is just a generalization of the previous theorems to the case of a
game of strategy. From Theorem 2, we know that the players behave rationally (avoid ex post
arbitrage) if and only if there is a supporting probability distribution that assigns non-negative
expected value to every gamble accepted by every player and assigns positive probability to the
event that occurs. The supporting probability distribution can be interpreted to represent the
commonly-held beliefs of the agents—i.e., the beliefs of a representative agent—notwithstanding
the distortions that may be introduced by state-dependent marginal utility for money. Now, if
41
the situation happens to be a game of strategy, and if the players have constant marginal utility
for money, and if they accept gambles which reveal their relative utilities for outcomes of the
game in the manner illustrated above, then the supporting probability distribution must be an
objective correlated equilibrium of the game defined by those utilities (Nau and McCardle 1990).
An objective correlated equilibrium is a generalized Nash equilibrium in which strategies of
different players are permitted (but not required) to be correlated19 (Aumann 1974, 1987), and the
game analyzed above (matching pennies) happens to have a unique correlated equilibrium that is
also the unique Nash equilibrium. So, once the players have revealed their utility functions in the
matching-pennies game via appropriate gambles, there is only one possible correlated
equilibrium distribution, and it represents the apparent common beliefs of the players. Although
the apparent beliefs of the players are thus restricted in this game, the no-arbitrage requirement
does not restrict the outcome that can occur, since all outcomes occur with positive probability in
some correlated equilibrium. In more general games, the only rational outcomes are those that
lie in the support of the set of correlated equilibria, and some individual or joint strategies may be
forbidden.20
What if the players do not have constant marginal utility for money? Then the same result still
holds, except that the correlated equilibrium distribution must be reinterpreted as a risk-neutral
distribution, not as the true probability distribution of any player. In this case, the true
probability distributions of the players form a subjective correlated equilibrium—i.e., an
equilibrium in which every player is maximizing her own expected utility, but the probability
distributions of different agents need not be mutually consistent (Nau 1995c). However, the
risk-neutral probability distributions of all agents (the products of their probabilities and relative
marginal utilities for money) must still be consistent. Hence, the Common Prior Assumption
applies to the players’ risk-neutral probabilities, not their true probabilities.
THEOREM 4 Game players are rational (avoid ex post arbitrage) if and only if (a) for
each player there is a probability distribution and utility function assigning non-negative
expected marginal utility to every gamble she accepts; (b) those probability distributions
constitute a subjective correlated equilibrium with respect to those utility functions; and
(c) there is a common prior risk neutral probability distribution assigning positive
probability to the observed outcome of the game.
Proof: Construct a matrix X whose rows are indexed by ijk, where i denotes a player and
j and k denote distinct strategies of that player, and whose columns are indexed by s,
where s denotes an outcome of the game—i.e., a joint strategy of all players. Let ui(s)
denote the utility payoff of player i in outcome s, let 1jk(s) be the indicator function for the
19
A correlated equilibrium can be implemented with a correlated randomization device, but the allowance of
correlation does not imply that such a device is being used, nor that the players are using randomized strategies at all.
Rather, the players’ moves may be uncertain and correlated only from the perspective of the outside observer.
Interestingly, it is possible for an outcome of a game to be “rationalizable” in the sense of Bernheim (1984) and
Pearce (1984) and yet not be rational in the sense of avoiding arbitrage. Conversely, in games with more than two
players, it is possible to have outcomes that are rational but not rationalizable, because they are supported only by
equilibria in correlated strategies. Examples are given in Nau and McCardle (1990).
20
42
event that player i plays strategy j, and let u(k, s-i) denote the utility payoff that player i
would have received by “defecting” to strategy k while all other players adhere to joint
strategy s. If the players are risk neutral—i.e., if they have constant marginal utility for
money—let the ijkth row of X be defined by xijk(s) = 1ij(s)(ui(s) – u(k, s-i)). This is the
payoff vector of a gamble that is acceptable to player i in the event that she plays strategy
j when she could have played k instead. By Theorem 0, either there is ex post arbitrage
in outcome s or else there is a probability distribution  such that X  0 and (s) > 0.
This is precisely the system of inequalities defining  as an objective correlated
equilibrium of the game defined by the payoff functions {ui(s)} (Aumann 1987, Nau and
McCardle 1990). In other words, either ex post arbitrage is possible is possible in
outcome s (i.e., strategy s is irrational, or “jointly incoherent”) or else s occurs with
positive probability in a correlated equilibrium of the game.21 If the players are not risk
neutral, let the ijkth row of X be defined by xijk(s) = 1ij(s)(ui(s) – u(k, s-i))/vi(s), where vi(s)
is the relative marginal utility for money of player i in outcome s. The system of
inequalities X  0 then defines an arbitrage-free equilibrium (Nau 1995c), which is a
subjective correlated equilibrium in which the players have common prior risk-neutral
probabilities.
Note that the imponderable mystery of game theory—the infinite regress of reciprocal
expectations of rationality—has been entirely finessed away. In the example, we said nothing
whatever about what Alice believed about what Bob believed about what Alice believed... Yet
the infinite regress is implicit in the criterion of no-arbitrage when it is applied to the players as a
group. If Alice behaves irrationally on her own (e.g., chooses a dominated alternative), that is an
arbitrage opportunity all by itself. If Bob bets on Alice to behave irrationally (e.g., chooses an
alternative that could only be rational for him if Alice behaved irrationally), then that too is an
arbitrage opportunity. (If Bob is wrong about Alice, you can collect a little from him, and if he is
right, you can collect a lot from Alice and more than cover your loss to Bob.) And if Alice bets
on Bob to bet on Alice to behave irrationally, that too is an arbitrage opportunity, and so on up
the ladder. The simple requirement of “no ex post arbitrage” automatically extrapolates the
sequence to infinity—and a little bit more. Meanwhile, the Common Prior Assumption has
mutated from virulent form into a harmless one: the requirement of common prior risk neutral
21
For example, in a generic 2x2 game with payoff matrix:
L
R
T
a, a
b, b
B
c, c
d, d
the matrix X has the following structure:
TL
TR
BL
BR
1TB
a–c
b–d
1BT
c–a
d–b
2LR
a – b
c – d
2RL
b – a
d – c
The first row is the payoff vector of the gamble that player 1 will accept in the event that she plays Top when Bottom
is available, etc. Here, utility functions have been attributed to the players in the specification of the “true” rules of
the game, following common practice. But given only knowledge of a matrix X of gambles that the players are
willing to accept, conditional on their own strategy choices, it is always possible to construct utility functions for
them that agree with X in the sense defined here. The matrix X encodes the “revealed” rules of the game.
43
probabilities is merely the condition for a competitive equilibrium in the market for contingent
claims on outcomes of the game, and it is the natural result of verbal behavior (money-backed
communication) among the players.
Example 5: learning from experience
Suppose that it is now 10:00pm, and as Alice is preparing to go to bed she sets the alarm on her
clock radio to come on at 7:00am, knowing that at 7:01am Bob (the radio weatherman) will
predict whether it is going to rain or not.
Alice at 10:00pm: I think the chance of rain is tomorrow is exactly 40%, and I also think
that if it’s going to rain, there’s an 80% chance that Bob will predict rain, and if it’s not
going to rain, there’s only a 30% chance that he will predict rain.
In other words, Alice is willing to accept any of the following bets:
Payoff for bet #1
Payoff for bet #2
Payoff for bet #3
Rain,
Prediction
$3
$1
$0
Rain,
No prediction
$3
-$4
$0
No rain,
Prediction
-$2
$0
-$7
No rain,
No prediction
-$2
$0
$3
(For simplicity, we assume in this example that Alice is able to state an exact chance for every
event—i.e., a rate at which she would bet indifferently on or against. Hence the bets in the table
above may multiplied by arbitrary positive or negative scaling factors.)
Here, Alice’s apparent “prior” probability of rain is 40%, and she has announced a likelihood
function that reflects her beliefs about the reliability of Bob’s forecast. It is unnecessary to ask
about her “posterior” probability of rain given Bob’s forecast, because it is implicit in what she
has already told us. For, the bets in the table above can be scaled and added up as follows:
Payoff for bet #1 (x0.96)
Payoff for bet #2 (x0.72)
Payoff for bet #3 (x0.64)
Total #1
Rain,
Prediction
$2.88
$0.72
$0
$3.60
Rain,
No prediction
$2.88
-$2.88
$0
$0
No rain,
Prediction
-$1.92
$0
-$4.48
-$6.40
No rain,
No prediction
-$1.92
$0
$1.92
$0
The total payoff is equivalent to betting as through there were a 64% chance of rain given a
prediction of rain, so Alice’s apparent posterior probability of rain given a prediction of rain is
64%. Of course, this is the same result we would have gotten by applying Bayes’ Theorem, but
Bayes Theorem—along with all the other “laws of probability” —is implicit in our calculus of
beliefs that is based on linear combinations of bets and avoidance of arbitrage (de Finetti 1937).
44
On the other hand, we can scale and combine the bets as follows:
Payoff for bet #1 (x0.56)
Payoff for bet #2 (x-1.68)
Payoff for bet #3 (x-0.16)
Total #2
Rain,
Prediction
$1.68
-$1.68
$0
$0
Rain,
No prediction
$1.68
$6.72
$0
$8.40
No rain,
Prediction
-$1.12
$0
$1.12
$0
No rain,
No prediction
-$1.12
$0
-$0.48
-$1.60
which is equivalent to betting as though there were a 16% chance of rain given a prediction of no
rain, which again is the posterior probability we would have gotten from Bayes Theorem.
To sum up the situation thus far, Alice is apparently expecting to learn from Bob’s prediction.
She would revise her prior probability of rain from 40% up to 64% if Bob were to predict rain,
and she would revise it down to 16% if he were to predict no rain, in a manner consistent with
Bayes Theorem. More precisely, if we were to ask her (at 10:00pm tonight) at what rate she
would bet on rain conditional on a prediction of rain or no rain, she would be irrational if she
gave answers other than 64% and 16%, respectively. For, suppose (hypothetically) that she made
the following additional statement:
Alice (hypothetically) at 10:00pm: I think the chance of rain given that Bob predicts rain
is 96%.
This would mean she would accept the following bet:
Payoff for bet #4
Rain,
Prediction
-$24
Rain,
No prediction
$0
No rain,
Prediction
$1
No rain,
No prediction
$0
...which could then be combined with the result of the previous bets:
Payoff for bet #4
Total #1 (from above)
Grand total
Rain,
Prediction
-$24.00
$3.60
-$21.40
Rain,
No prediction
$0
$0
$0
No rain,
Prediction
$1.00
-$6.40
-$5.40
No rain,
No prediction
$0
$0
$0
...earning a riskless profit from Alice in the event that Bob predicts rain. With suitable scaling of
bets, the same trick could be played for any asserted chance of rain, given a prediction of rain,
other than 64%. We conclude that Alice would be irrational in the event that Bob predicts rain if
her (hypothetical) posterior probability in that event were anything other than the value
calculated from her prior probability and her likelihood function according to Bayes Theorem.
45
Rational learning thus seemingly requires probabilities to be updated in accordance with Bayes
Theorem. More precisely, rational expected learning requires probabilities to be updated in
accordance with Bayes Theorem: it is still 10:00pm on the night before, and no actual learning
has yet taken place.
Now let us fast-forward to the next morning. At 7:00am the radio comes on, at 7:01 Bob predicts
that it will rain, and suppose that immediately afterward Alice says the following:
Alice (actually) at 7:02am: I think the chance of rain today is 96%.
Is she now irrational? Not at all! The fact that her betting rate today, given that Bob has
predicted rain, differs from the “posterior” probability calculated the night before does not create
any opportunity for arbitrage. An observer could not have predicted at 10:00pm last night what
Alice would do at 7:02am today, and so he could not have placed a combination of bets that
carried no risk of loss for him. In fact, Alice is free to bet on rain at any rate whatever following
the announcement of Bob’s forecast, unconstrained by the beliefs she expressed the night before.
It seems that rational learning over time need not follow Bayes Theorem after all, as has been
argued by Hacking (1967), Goldstein, (1983, 1985), and the Austrian-school economists.
But this is not quite all we can say about learning over time. Suppose that on the night before,
we ask Alice what she believes her betting rate on rain will be the next morning—say,
immediately prior to hearing Bob’s forecast—and that she makes the following statement:
Alice (additionally) at 10:00pm: I think there’s a 2/3 chance that when I get up in the
morning I will say the chance of rain is 15% and a 1/3 chance that I will say the chance of
rain is 90%
(Perhaps she will look out the window at 7:00am to see whether the sky looks threatening, or
perhaps she will lie awake calculating the odds of rain more carefully between now and then.)
This additional statement enlarges the environment to include Alice’s betting rate in the morning
as another event about which there is uncertainty tonight. As such, it does establish a connection
between Alice’s betting rate tonight and her betting rate tomorrow morning. To see this, note
that Alice is now accepting the following additional bets:
Payoff for bet #4
Payoff for bet #5
Payoff for bet #6
Payoff for bet #7
15%,
Rain
$.33
-$.33
$.85
$.00
15%,
No rain
$.33
-$.33
-$.15
$.00
90%,
Rain
-$.67
$.67
$.00
$.10
90%,
No rain
-$.67
$.67
$.00
-$.90
(The event label “15%, Rain” means that Alice’s probability of rain in the morning is 15% and it
subsequently rains, etc. Bob’s predictions have been suppressed in this table, since none of the
four new bets depends on his prediction.) Bets #4 and #5 are bets on the value of Alice’s betting
rate in the morning. Thus, for example, Alice will accept a bet (tonight) in which she will win
46
$0.33 if her betting rate tomorrow is 15% and lose $0.67 otherwise. Bets #6 and #7 are bets on
rain given the value of Alice’s betting rate tomorrow, which are acceptable by definition of the
“betting rate” events: in the event that her betting rate tomorrow is 15%, she will by definition
accept a bet in which she wins $0.85 if it rains and loses $0.15 if it doesn’t rain, and so on. The
new bets can now be combined as follows:
Payoff for bet #4 (x0)
Payoff for bet #5 (x0.75)
Payoff for bet #6 (x1)
Payoff for bet #7 (x1)
Total
15%,
Rain
$.00
-$.25
$.85
$.00
$.60
15%,
No rain
$.00
-$.25
-$.15
$.00
-$.40
90%,
Rain
$.00
$.50
$.00
$.10
$.60
90%,
No rain
$.00
$.50
$.00
-$.90
-$.40
...which is the same as betting on rain (tonight) at a rate of 40%. Fortunately, this is the same
rate at which Alice previously stated she would bet on the occurrence of rain, otherwise an
arbitrage opportunity would exist. Not coincidentally, 40% is the expected value of Alice’s
betting rate in the morning, according to the beliefs about tomorrow’s betting rate that she
expresses tonight—that is 40% = (2/3)15% + (1/3)90%. Of course there is a theorem which
says this must be true, and it is just the theorem on the “prevision of a prevision” originally
proved by Goldstein (1983):
THEOREM 5: An agent’s beliefs about her future beliefs are rational (avoid arbitrage)
if and only if her current betting rate on an event equals the expected value of her future
betting rate.
In other words, rationality requires the agent to expect her beliefs to unfold over time according
to a martingale process. Indeed, this theorem is just a simplified version of the Harrison-Kreps
(1979) theorem on arbitrage and martingales in securities markets. Here the agent imagines her
personal probability of rain to be a stochastic process with some degree of volatility over time,
analogous to the stochastic process describing a stock price.
The preceding theorem requires the agent’s beliefs to evolve according Bayes theorem only in
expectation. In the example, if Alice tonight believes that Alice tomorrow will believe the
chance of rain to be either 15% or 90% in the instant before she hears Bob’s forecast, and if her
likelihood function meanwhile remains the same, then according to Bayes’ Theorem, her
corresponding posterior probability for rain given a prediction of rain will either be 96% or 32%.
As of 10:00pm on the night before, these two values appear equally probable,22 yielding an
It might naïvely be expected that the different values of 32% and 96% for Alice’s posterior probability of rain,
given a prediction of rain, would occur with probabilities 2/3 and 1/3 respectively, but this is not the case. With
probability 2/3, Alice’s prior probability at 7:00am will be 15%, in which case, based on her likelihood function, the
marginal probability of a prediction of rain will be 37.5%, so the joint probability of “prior=15% & prediction=rain”
is (2/3)37.5% = 25% (as of 10pm on the night before). With probability 1/3, Alice’s prior probability at 7:00am will
be 90%, in which case the marginal probability of of a prediction of rain will be 75%, so the joint probability of
“prior=90% & prediction=rain” is (1/3)75% = 25%. Hence, conditioning on a prediction of rain, it is equally likely
22
47
expected posterior probability of 64%, in agreement with what was calculated earlier. Bayes’
Theorem describes actual learning over time only in the degenerate case where the agent feels
that there is “zero volatility” in her beliefs, so that the probability distribution for her future
beliefs is a point mass. (In such a case, the agent would say “there is a 100% chance that my
betting rate tomorrow will be equal to p” for some value of p, and if her betting rate turns out to
be anything else, she is exposed to arbitrage.) Since there is no reason why, in general, beliefs
should not vary over time for unexplained reasons—due to the fruits of deeper introspection, the
receipt of totally unanticipated information, or perhaps cosmic rays colliding with neurons in the
brain—Bayes theorem is not a general model of learning from experience.23
The usual objection to this “proof” that agents do not actually learn according to Bayes’ Theorem
is to argue that the original model was merely incomplete (e.g., Howson and Urbach 1989). If
we had included more informational events in the original description of the environment—say,
the result of looking out the window at 7:00am or an account of what happened during the
night—we would have been able to predict the evolution of beliefs over time in actuality, not
merely in expectation. But the rejoinder to this objection is that no matter how many
informational events are anticipated, no matter how finely the states of nature are partitioned,
there will still be some events that are not discriminated ex ante but are relevant to the evolution
of beliefs ex post. By allowing room in the model for unexplained volatility in beliefs, we
thereby include a proxy for unforeseeable informational events. Even if the agent cannot
enumerate or describe the unforeseeable (or otherwise unmodeled) events, she may yet be able to
estimate the amount of volatility in her beliefs that she expects to occur over time.
Example 6. Allais’ paradox
The example of an empirical violation of the axioms of expected utility first concocted by Allais
(1951) has inspired many of the theories of non-expected utility developed over the last two
decades. An analysis of this example from the perspective of arbitrage choice theory will help to
illustrate how ACT departs from both EU and non-EU theory, in which preference is a behavioral
primitive. In a simpler version of the paradox, introduced by Tverky and Kahneman (1979), a
subject is presented with the following two pairs of alternatives.
A: 100% chance of $3000
B: 80% chance of $4000
A: 25% chance of $3000
B: 20% chance of $4000
that the prior is 15% or 90%, in which case it is also equally likely that the posterior is 32% or 96%, so the expected
posterior probability is (1/2)32% + (1/2)96% = 64%, in agreement with the original Bayes Theorem calculation.
In the example, Alice’s “prior” probability is imagined to change before she hears Bob’s forecast. There is no
reason why her likelihood function could not also change—e.g., Bob’s prediction might surprise her in some way,
triggering a larger or smaller revision in belief than she had anticipated. As Hacking (1967) observes: “The man’s
P(H|E) before learning E differs from his P(H|E) after learning E. Why not, he says: the change represents how I
have learned from E!”
23
48
(In all cases, the remaining probability mass leads to a payoff of $0.) The typical response
pattern is that A is preferred to B but B is preferred to A, in violation of the independence
axiom: most persons would rather have a sure gain than a risky gamble with a slightly higher
expected value, but they would maximize expected value when faced with two “long shots.”
This pattern of behavior does not necessarily violate the assumptions of arbitrage choice theory,
because from the perspective of ACT, the decision problem is ill-specified:

The choices are completely hypothetical: none of the alternatives is actually available.

They occur in different hypothetical worlds: in the world where you choose between A and B
a sure gain is possible, while in the world where you choose between A and B it is not.

The relations between the events are ambiguous: it is not clear how the winning event in
alternative B is related to the winning events in alternatives A  or B.
In order to recast this example in the framework of ACT (or otherwise turn the independenceaxiom-violator into a bona fide money pump), the choices must be real, they must be forced into
the same world, and the relations between events must be made explicit. The following decision
tree shows the usual way in which such details are added to the problem description (e.g.,
Seidenfeld 1988, Machina 1989) in order to materially exploit preference patterns that violate the
independence axiom:
$3000
Safe
25%
2
80%
1
Risky
75%
$0
$4000
3
20%
$0
Here, the choice between A and B is the choice between “safe” and “risky” if a decision is to be
made only upon reaching node 2, whereas the choice between A and B is the choice between
“safe” and “risky” if an irrevocable commitment must be made at node 1. In this refinement of
the problem, the winning event in alternative B is a proper subset of the winning event in
alternative A, and the winning event in B is also conditionally the same as the winning event in
B, given that node 2 is reached. Savage (1954) pointed out that when the relations among the
events are made explicit in this fashion, the paradox vanishes for many persons, including
himself (and myself): the commitment that should be made at node 1 is determined by imagining
the action that would be taken at node 2 if it should be reached, thereby ensuring consistency.
Or, alternatively, the agent may “resolutely” choose to honor at node 2 whatever commitment
was made at node 1 (McClennen 1990).
49
But suppose that even when presented with this picture, an agent insists on preferring A to B and
B to A. At this point it is necessary to be explicit about which “world” the agent is in and at
which point in time she is situated in it. At most one of these choices is real at any given
moment, and it is the only one that counts. For, suppose that the agent is at node 1 in possession
of alternative A in the form of a lottery ticket entitling her to the proceeds of the “safe” branch in
the event that node 2 is reached. Then for her to say that she prefers B to A means that she
would be willing to pay some $ > 0 to immediately exchange her “safe” ticket for a “risky”
ticket, if permitted to do so. But what does it mean, at node 1, for her to also say that she prefers
A to B? Evidently this assertion must be interpreted merely as a prediction of what she would at
some point in the future that may or may not be reached, namely, she predicts that upon
reaching node 2 she would be willing to pay $ to switch from a “risky” to a “safe” ticket, if
permitted to do so at that point. If this entire sequence of exchanges and events indeed
materializes, then at node 2 she will be back in her original position except poorer by $2.
Meanwhile, if she does not reach node 2, both tickets will be worthless, and she will still be
poorer by $. Does this pattern of behavior constitute ex post arbitrage? Not necessarily! A
agent’s prediction of her future behavior realistically should allow for some uncertainty, which
could be quantified in terms of bets in the fashion of the example in the preceding section. But in
order for a clear arbitrage opportunity to be created, the agent must not merely predict at node 1
that she will switch back from “risky” to “safe” upon reaching node 2, she must say at node 1
that she is 100% certain she will do so at node 2, which is equivalent to an irrevocable
commitment to do so. (It means she is willing to suffer an arbitrary financial penalty for doing
otherwise.) But an irrevocable commitment at node 1 to exchange “risky” for “safe” at node 2 is
then equivalent to immediately switching from B back to A. It is doubtful that anyone would
pay $ to exchange A for B and, in the next breath, pay another $ to undo the exchange, and so
the paradox collapses.
The key issues are (a) whether an irrevocable commitment to “safe” or “risky” is or is not
required at node 1, and if it is not, then (b) what is the meaning of a “preference” held at node 1
between alternatives that are to be faced at node 2. In world she really inhabits, either the agent
will have the opportunity make her choice at node 2 after some interval of time (measured by the
realization of foreseen and unforeseen events) has passed, or else she must choose immediately.
If she doesn’t have to choose immediately, then she might as well wait and see what happens. In
that case, the only choice that matters is between A and B, and it will be made, if necessary, at
node 2. If she is currently at node 1, anything that she says now concerning her behavior at node
2 is merely a prediction with some attendant uncertainty. (In some situations it is necessary for
the agent to predict her own future choices in order to resolve other choices that must be made
immediately. But even in such situations, as we saw in example of the preceding section, it is
permissible for an agent to be somewhat uncertain about her future behavior and to change her
mind with the passage of time.) Meanwhile, if the agent does have to choose immediately, the
only choice that matters is between A and B. If she chooses B over A believing that those are
her only options, but later an opportunity to exchange B for A unexpectedly materializes at node
2—well, that is another day!
In summary:
50

Standard choice theory requires agents to hold preferences among objects in hypothetical
(and often counterfactual) worlds that may be mutually exclusive, or only ambiguously
connected, or frozen at different moments in time. The independence axiom then forces
consistency across such preferences.

Allais’ paradox exploits the fact that preferences may refer to different worlds. (The
objective-probability version also exploits the ambiguity concerning the relations of events
between worlds.) But most agents construct their preferences by focusing on salient features
of the world they imagine they are in, and a world in which a sure gain is possible is
perceived differently from a world in which winning is a long shot in any case (Shafer 1986).
Hence, their preferences in different worlds may violate the independence axiom.

In ACT, by comparison, there are no hypothetical or counterfactual choices: there is only
one world, and every act of verbal or physical behavior has material implications in that
world. Moreover, consistency is forced only on behavior that occurs at the present moment
in the one world. When agents are called upon to make predictions about their future
behavior, they are allowed some uncertainty, and such uncertainty leaves room for them to
change their minds over time without violating the no-ex-post-arbitrage axiom.
3.3 On the realism and generality of the modeling assumptions
The modeling assumptions of arbitrage choice theory are simpler and more operational than than
those of the standard theory, but in some ways they may also appear to be less realistic and less
general in their scope. For one thing, money plays a prominent role, whereas rational choice
theorizing is often aimed at nonmonetary behavior and noneconomic institutions (e.g., voting).
And not only do interactions among agents in our models typically involve money, but they take
place in a market where it is possible to bet on practically anything, including one’s own
behavior. Outside of stock exchanges, insurance contracts, casinos, office pools, and state
lotteries, most individuals do not think of themselves as bettors, and bets that explicitly depend
on one’s own choice behavior are uncommon (though not unheard of). The standard theory, by
comparison, builds on a foundation of preferences (which everyone has to some degree) and
consequences (which can be anything whatever). At first blush, the latter approach seems
flexible, broadly applicable, and perhaps even uncontroversial. However, it later must pile on
assumptions about imaginary acts, common knowledge, common belief, and equilibrium
selection that go far beyond what has been assumed in the examples of this section. If such
additional assumptions were indeed realistic, the market transactions in the arbitrage models
would not seem at all fanciful: everyone would already know the gambles that were acceptable
to everyone else, and more besides.
The alternative theory of choice presented here is idealized in its own way, but nevertheless it is
grounded in physical processes of measurement and communication that actually enable us to
quantify the choice behavior of human agents with some degree of numerical precision. In the
familiar institutions that surround us, money usually changes hands when subjective beliefs and
51
values must be articulated in quantitative terms that are both precise and credible. Moreover, any
exchange of money for goods or services whose reliability or satisfactoriness or future value is
uncertain is effectively a gamble, and in this sense everyone gambles all the time. The markets in
which such “gambles” take place not only serve to allocate resources efficiently between buyers
and sellers, but also to disseminate information and define the numerical parameters of common
knowledge. Indeed, the common knowledge of a public market is the intuitive ideal against
which other, more abstract, definitions of common knowledge are compared. Last but not least,
contracts involving monetary payments are quite often used to modify the “rules” of games
against nature or games of strategy: agents use contingent contracts to hedge their risks or attach
incentives or penalities to the actions of other agents. Monetary transactions and market
institutions are therefore important objects of study, even if they are not the only such objects.
(Indeed, the line between market and non-market institutions is rather blurry nowadays.) Insofar
as money and markets play such a fundamental role in quantifying beliefs and values, defining
what is common knowledge, and writing the rules of the games we play, it is reasonable to
include them as features of the environment in a quantitative theory of choice, at least as a point
of departure. Even where real money does not change hands or a real market does not exist in
the problem under investigation, they may still serve as useful metaphors for other media of
communication and exchange.
It might be expected that, by erecting the theory on a foundation of money and highly idealized
markets, strong results would be obtained that, alas, would lose precision when extrapolated to
imperfect markets or non-economic settings. It is interesting, then, that the results we have
obtained are more modest in some ways than the results claimed under the banners of the
standard theory, and in other ways orthogonal to it.
3.4 Summary
The preceding examples illustrate that the assumptions of arbitrage choice theory lead to
conclusions about rational behavior that resemble those of the standard theory in many respects.
First, it is sufficient (but not necessary) for every agent to behave individually in a manner that
maximizes expected utility with respect to some probability distribution and utility function.
(Non-expected-utility preferences are allowed as long as they imply that more wealth is always
preferred to less—see Nau 1999 for an example.) Second, the agents jointly must behave as if
implementing a competitive equilibrium in a market or a strategic equilibrium in a game. Third,
where uncertainty is involved, there must exist a “common prior” probability distribution. And
fourth, where agents anticipate the receipt of information, they must expect their beliefs to be
updated according to Bayes Theorem. However, these results also differ from those of the
standard theory in some important respects, namely:

Probability distributions and utility functions need not be uniquely determined, i.e., belief and
preference orderings of agents need not be complete, nor do they need to be separable across
mutually exclusive events.

Equilibria need not be uniquely determined by initial conditions.
52

“True” probabilities and utilities, to the extent that they exist, need not be publicly observed
or commonly known: they are generally inseparable.

Utility functions need not be state-independent.

The uncertain moves of different agents need not be regarded as probabilistically
independent.

Bayes’ Theorem need not describe the actual evolution of beliefs over time.

Common prior probabilities are risk neutral probabilities (products of probabilities and
relative marginal utilities for money) rather than true probabilities.
Among these departures from the standard theory, the reinterpretation of the Common Prior
Assumption is the most radical: it is literally at odds with the most popular solution concepts in
game theory and information economics. It calls attention to a potentially serious flaw in the
foundations of game theory and justifies much of the skepticism that has been directed at the
Common Prior Assumption over the years.
The other departures are all in the direction of weakening the standard theory: personal
probabilities and utilities are not uniquely determined, beliefs and values are confounded in the
eye of the beholder, equilibria are coarsened rather than refined, and the evolution of beliefs is
stochastic rather than deterministic. Much of this ground has been traveled before in models of
partially ordered preferences, state-dependent utility, correlated strategies, and temporally
coherent beliefs. Assumptions of the standard theory are weakened and, not surprisingly, weaker
conclusions follow. The arbitrage formulation scores points for parsimony and internal
cohesiveness—especially in its tight unification of decision analysis, game theory, and market
theory—but otherwise it appears to settle for rather low predictive power.
But there is one other very important respect in which arbitrage choice theory departs from the
standard theory, which was alluded to earlier: the arbitrage theory does not rest on assumptions
about the rationality of individuals. Its axioms of rational behavior (additivity and no ex post
arbitrage) apply to agents as a group: they carry no subscripts referring to specific individuals .
It is the group which ultimately behaves rationally or irrationally, and the agents as individuals
merely suffer guilt-by-association. The fact that the group is the unit of analysis for which
rationality is defined admits the possibility that the group is more than just an uneasy alliance of
individuals who are certain of and wholly absorbed in their own interests. We have already seen
that the beliefs and values of the group (its betting rates, prices, etc.) are typically more sharply
defined than those of any of its members, but that is only the tip of a larger iceberg, as the next
section will explore.
53
4.1 Incomplete models and other people’s brains
Standard choice theory assumes that a decision model is complete in every respect: the available
alternatives of agents and the possible states of nature are completely enumerated, the
consequences for every agent in every event are completely specified in terms of their relevant
attributes, and the preferences of the agents with respect to those consequences are (among other
things) completely ordered. As a result, it is possible to draw a diagram such as the following for
a choice problem faced by a typical agent:
Figure 1:
Attribute
Consequence
u11
p
State
v11
u12
a1
Alternative
1-p
v12
u21
p
a2
v21
u22
1-p
v22
In this example, the agent chooses one of two possible alternatives, then one of two possible
states of nature obtains, and the agent receives a consequence having two attributes. Because the
agent is also assumed to have completely ordered preferences over a much larger set of acts, we
can infer the existence of unique numerical probabilities of events (here indicated by p and 1-p)
and unique (up to affine transformations) utilities of consequences. Under suitable additional
axioms on preferences (Keeney and Raiffa 1976), the utilities of consequences can be further
decomposed into functions of the utilities of their attributes. For example, it might be the case
here that the utility of a consequence is an additive function of the utilities of its attributes, so
that, after suitable scaling and weighting, the utility of consequence 1 on the upper branch is u11
+ v11, where u11 is the utility of the level of attribute 1 and v11 is the utility of the level of attribute
2, and so on. Thus, we can compute the agent’s expected utilities for alternatives a1 and a2:
54
EU(a1) = p(u11 + v11) + (1-p)(u12 + v12)
EU(a2) = p(u21 + v21) + (1-p)(u22 + v22)
The optimal alternative is the one with the higher expected utility, and it is uniquely determined
unless a tie occurs, in which case there is a set of optimal alternatives among which the agent is
precisely indifferent. This is a satisfactory state of affairs for the agent, who knows exactly what
she ought to do and why she ought to do it, and it is also satisfactory for the theorist, who can
predict exactly what the agent will do and why she will do it. And if there are many agents, the
omniscience of the theorist is merely scaled up: she can predict what everyone will do and why
they will do it. A complicated economic or social problem has been reduced to a numerical
model with a small number of parameters that easily fits inside one person’s head (perhaps aided
by a computer).
Of course, a “small world” model such as the one shown above is often merely an idealization of
a more complicated “grand world” situation. More realistically, the agent’s beliefs and values
may be only incompletely ordered, in which case probabilities and utilities are typically
represented by intervals rather than point values:
Figure 2:
Attribute
Consequence
[u11 , U11]
[p, P]
State
[v11 , V11]
[u12, U12]
a1
Alternative
[1-P, 1-p]
[v12, V12]
[u21, U21]
[p, P]
a2
[v21 , V21]
[u22 , U22]
[1-P, 1-p]
[v22 , V22]
Here, p and P denote lower and upper bounds, respectively, on the probability of the event, and
uij and Uij denote lower and upper bounds, respectively, on the utility of attribute 1 on the ijth
55
branch, and so on. We can no longer compute an exact expected utility for each alternative, but
we can compute lower and upper expected utilities (denoted eu and EU, respectively):
eu(a1) = min {x(y11 + z11) + (1-x)(y12 + z12)}
EU(a1) = max {x(y11 + z11) + (1-x)(y12 + z12)}
where the minimum and maximum are taken over all x, y11, z11, etc., satisfying x  [p, P], y11 
[u11, U11], z11 [v11, V11], etc. If it turns out that EU(ai) < eu(aj)—i.e., if the intervals of expected
utilities are disjoint—then we may still conclude that there is a unique optimal solution. But, in
general, it is possible to have multiple “potentially optimal” solutions when there is overlap
among the expected utility intervals for different alternatives. The optimal solution is now setvalued, as are the parameters of the model. (Models of this kind have been studied by Smith
1961, Aumann 1962, Giron and Rios 1979, Rios Insua 1990, Walley 1991, Nau 1989 & 1992a,
Seidenfeld et al. 1998, among others.) In the worst case, we may not be able to pin the optimal
solution down very precisely, but we can often at least narrow the range of alternatives that ought
to be considered.
The model of Figure 2 (with incomplete preferences) is undoubtedly more realistic than the
model of Figure 1 (with complete preferences), but it is somewhat unsatisfying for the agent and
perhaps even more unsatisfying for the theorist. The agent merely fails to get a unique
recommended solution from the model, whereas the theorist’s entire house of cards begins to
wobble: the infinite regress, the common knowledge assumptions, the refinements of equilibria,
all become hard to sustain in the face of indeterminacy at the agent level.
56
But the true situation is even worse: realistically, most models are not only preferentially
incomplete, they are also dynamically and/or structurally incomplete:
Attribute
Figure 3:
Consequence
State
Alternative
The attributes of consequences, states of nature, and available alternatives may not be fully
enumerated, and the timing of choices—even the necessity of making choices—may be
indeterminate, as bounded-rationality theorists emphasize.24 This is the situation normally
encountered at the first stage of decision analysis: the tree has not yet been drawn and the
decision maker has not yet thought of everything she might do, everything that might happen,
and all the consequences that might befall her—let alone quantify her beliefs and values. In such
a situation, the set of alternatives that are even potentially optimal may be indeterminate. This is
not to say that analysis might not be helpful—indeed, the very lack of clarity that a decision
maker typically experiences upon facing an unfamiliar problem is the raison d’être for
professional decision analysts. Structured techniques of “value-focused thinking” (Keeney 1992)
and “strategy generation” (Howard 1988) can be used to articulate values and create strategic
alternatives, the views of other stakeholders can be solicited to provide different perspectives and
new ideas, prototype models can be tested to identify the most important tradeoffs or
Simon (1947, chapter IV) comments: “The task of decision involves three steps: (1) the listing of all the
alternative strategies; (2) the determination of all the consequences that follow upon each of these strategies; (3) the
comparative evaluation of these sets of consequences. The word ‘all’ is used advisedly. It is obviously impossible
for the individual to know all his alternatives or all their consequences, and this impossibility is a very important
departure of actual behavior from the model of objective rationality.”
24
57
uncertainties, and finally the relevant probabilities and utilities can be assessed. Eventually a
model like the one in Figure 1 may be constructed, and it is hoped to be a “requisite” model of
the real-world situation.
What is significant about such prescriptive applications of rational choice theory to real decisions
is that most of the effort typically goes into completing the model rather than solving a model that
is already complete, and the process of completing the model usually involves other people’s
brains.25 Someone facing a novel decision often begins by asking the advice of colleagues and
then proceeds to get help from experts in the problem area and/or from the marketplace. For
example, a person requiring a new home appliance does not undertake an elaborate decision
analysis to determine how much she ought to pay for it, wracking her own brain to assess the
expected utility of the uncertain stream of benefits produced by owning the appliance and then
inverting it to obtain the corresponding certainty equivalent in dollars. She just goes to a local
store and pays the competitive market price, knowing that it embodies the uncertainty
assessments and value tradeoffs made by millions of other consumers interacting with scores of
producers and retailers.26 In the spectrum from cheap to expensive models, she focuses on a
price range and brand image that is appropriate for someone in her demographic cohort, an
identity that has been impressed on her by countless TV advertisements. If she undertakes any
analysis at all, it may be to ask her friends for recommendations, read consumer publications for
product ratings, and/or compare models and prices at a few stores, which further enlarges the
pool of brains she draws upon. In the process, she will discover a few things about the product
category that she was unaware she didn’t know. Her own private knowledge and idiosyncratic
values will play some role in her eventual “choice” between model X and model Y—which is her
own small contribution to the collective wisdom of the market—but by then most of the
important work of creating and pricing the alternatives and focusing her attention will have been
done by others.
If the decision problem arises in an organizational setting, standard administrative procedures
may be followed to assign personnel, gather information, construct and evaluate alternatives,
build consensus, and generate action. For example, a strategy team may be formed; committee
meetings, focus groups, and decision conferences may be held; consultants may be brought in;
and dialogues among different levels of management may be conducted. Higher levels of
management are often occupied by individuals who scaled the corporate ladder precisely by
virtue of their knack for harnessing the abilities of subordinates and providing leadership in
ambiguous situations (Isenberg 1984, March and Simon 1993, March 1994). Eventually a
decision is made and an individual may assume responsibility, but by then many individuals have
contributed to the process, and it is possible that no one has a clear view of the entire problem—
At a commencement address in 1992, J.B. Fuqua remarked: “Perhaps the most important thing I learned early in
my business career was that I would not attain my personal or business goals unless I used OPM and OPB—other
people’s money and other people’s brains.” Much of economic theory is concerned with the use of OPM. Rather
less attention has been paid to the use of OPB, although recent work in complexity theory (e.g., Lane et al. 1996)
emphasizes the role of “networks of generative relationships” in business decision making.
25
26
Significantly, the market also strategically decouples the consumer and the store owner, so they need not haggle
over the price. Market-based communication of beliefs and values often has this effect of eliminating the need to
play games.
58
or even the same view as anyone else. As on a professional sports team, there are different views
from the playing field, the coaching box, and the front office, and all must contribute in order for
the team to succeed.27
Another aspect of model incompleteness, which was illustrated in Example 5, is that the mere
passage of time often lends detail to a decision model because unforeseen things happen, as the
Austrian economists and complexity theorists emphasize. Nature or a human competitor makes
an unexpected move, objectives wax or wane in perceived importance, opportunities that were
not previously envisioned suddenly present themselves, while alternatives that were believed to
be available sometimes turn out to be illusory. Indeed, the subjective experience of time is
measured by the accumulation of small surprises: if nothing that you did not precisely foresee
has happened lately, time has stood still for you. For this reason, obtaining the most “perfect”
information that can be envisioned today may not be equivalent to obtaining an option to put off
a decision until tomorrow, contrary to textbook decision analysis.
The preceding description of the decision making process suggests that all decisions are to some
extent group decisions, and they are acts of creation and discovery, not merely acts of selection.
This view is compatible with arbitrage choice theory—and incompatible with the standard
theory. The standard theory assumes that every agent begins with a complete mental model of
the problem, and such models are often required to be mutually consistent or even objectively
correct. Every agent is an island of rationality with no reason to visit the islands of other agents
except to get more of what she already knows she wants through trade or contests of strategy.28
In the arbitrage theory, by comparison, the rational actor is naturally a group. An individual
typically perceives only part of the problem: different agents know the relative values of different
objects, the relative likelihoods of different events, the ranges of alternatives that are appropriate
in different situations—and no one can anticipate everything that might happen tomorrow. Like
the proverbial blind man holding one limb of an elephant, an individual has opinions about some
things but not necessarily everything, and interactions with others may provide a more complete
picture. In the end, “rationality” gets enforced at the point where inconsistencies within or
between agents become exploitable, and exploitation usually occurs against groups rather than
individuals.29
27
Of course, through experience, individuals can acquire the capability to solve various kinds complex problems
single-handedly, and we then call them “experts.” But true experts usually find it hard to externalize the source of
their expertise, which resides mainly in the ability to recognize patterns and instinctively generate appropriate
responses. Hence the behavior of the expert cannot be easily explained or replicated by an analytic model.
Even when the agent seeks to acquire information from other agents, she is assumed to “know what she doesn’t
know.” That is, she is assumed to be able to foresee all her possible future states of information, and the process of
information-gathering merely instantiates one of these already-foreseen states. Similarly, when she hires specialists
to act on her behalf, she is assumed to be able to foresee the possible consequences of their actions and to quantify
the probabilities with which those consequences will obtain—she takes nothing “on faith.”
28
29
Empirical research on group decision making shows that when individuals make judgments about uncertain
quantities and later meet to form a group consensus judgment, the accuracy of the group judgment usually is greater
that of a typical individual’s judgment but less than that of the mean of the individual judgments. (Hill 1982, Kerr et
al. 1996, Gigone and Hastie 1997) In other words, it is usually better just to average the individual judgments rather
than hold a face-to-face meeting, apparently because the latter may lead to “group-think,” “risky shifts,”
59
4.2 The limits of theory
Of course, to be applied quantitatively, the arbitrage theory of rational choice requires a certain
degree of completeness: events must be well-defined and contingent trades must be possible, or
at least imaginable, with respect to those events. As such, the theory recognizes its own limits.
Where events are not well-defined and where contingent trades are hard even to imagine, we
should not expect to be able to predict or prescribe choice behavior in exact, quantitative terms.
Even there, the theory is not altogether silent: it suggests that completing the model may be the
most important task at hand and that interactive decision processes (or “networks of generative
relationships” to use the phrase of Lane et al. 1996) may facilitate that task. But a sobering
implication emerges, namely that if the agents typically do not see the whole picture, then neither
does the theorist. If the purpose of interactions among agents is to construct, between them, a
more complete model of a situation that is too complex for any individual to fully comprehend,
then the theorist cannot hope to do that job for them single-handedly. No single mode of analysis
or theoretical paradigm can expect to yield the last word in explaining a social or economic
phenomenon. This is not an argument for relativism in science, but merely a reflection on the
reason why universalist theories and quantitative models have thus far proved less useful in the
social sciences than in the natural sciences. Models yield accurate predictions only if their
complexity is commensurate with that of the phenomena they are intended to represent, which is
easier to achieve with physical systems (where we look into the eyepiece of the microscope) than
with social and economic systems (where we look up from the other end). If the complexity of
the system is very much greater than that of the observer and her model, she cannot expect to
understand every detail by herself—but the scientific establishment to which she contributes will
gradually assemble a more complete picture through its network of institutions, its competition
among paradigms, and its accumulation of technology and literature.
From this perspective, let us now return briefly to some of the questions raised earlier. First,
recall Elster’s (1986) assertion that a group such as a family does not have ‘its’ own goals and
beliefs. The concepts of goals and beliefs are operational only to the extent that they are
reflected in decisions or other observable behavior. The family often ‘decides’ as a unit, and in
that case, whose goals and beliefs are on view? The family may be composed of discrete
individuals—man, woman, child, and dog—but others contribute as well to its decisions:
financial and spiritual advisers, insurance and travel agents, architects, interior decorators, friends
and relatives, plus all the forces of commercialism, conformity, and historical contingency. The
family’s decision is therefore more than just a compromise or a strategic equilibrium among
goals and beliefs that were already present and well-formed in the heads of its nominal members.
Rather, it is an act of construction to which many contribute, whose precise outline none may
fully anticipate, and whose final causes perhaps none can fully explain. Of course, the family
does not possess the same unitary consciousness of its actions that we would attribute to an
overconfidence, underutilization of individual expertise, and so on . Much of this literature is based on laboratory
experiments in which individuals and groups solve artificial problems with a high degree of completeness, for which
there are correct answers against which performance can be objectively measured. The theory outlined in this paper
suggests that the benefits of group interaction will be most evident in decision problems that are characterized by a
low degree of completeness and/or which occur in competitive environments where inefficient behavior by groups is
subject to exploitation. But either a group consensus judgment or an average of individual judgments qualifies as a
group process for our purposes.
60
individual, but that is beside the point. We could ask the family to explain ‘its’ reasons for
taking a decision, and we would get an answer as the result of another act of construction. The
reasons offered might leave some things unexplained or ambiguous, but the same would be true
if the subject were an individual. This view of the family is not entirely incompatible with the
operational spirit of methodological individualism as defined by Watkins (1957), but it suggests
that the decision-making units that are the focus of the analysis need not always be individuals.
It does, however, imply that the companion concept of purposive action is too strong. The
decision-making units, whether or not they are individuals, generally will not be able to
completely and correctly articulate the causes and effects of their own behavior: that is why they
must rely on others.
Next, does a firm have a utility function? Insofar as it is a rational actor—which is to say, insofar
as it is a decision-making unit whose behavior does not give rise to either intra-firm or inter-firm
arbitrage—then probably it does act as if maximizing expected utility, subject to the caveats
noted above. But does this observation have practical implications for decision modeling?
Probably not—because no individual in the firm is the keeper of the utility function, and the most
important strategic decisions of the firm involve creating new alternatives and responding to
environmental change, not merely choosing consistently among fixed alternatives in a stable
environment. Hence, a decision analyst armed with a mythical corporate utility function cannot,
even in principle, do the work of the firm’s executives, managers, and administrative procedures.
Analysis and formal models may sharpen the abilities of corporate decision makers, but cannot
fully represent their function. There is no mathematical substitute for executive intuition and
leadership.
Why do individuals behave badly in many behavioral experiments? The tasks and the
surroundings are often unfamilar and only superficially simple, and the subjects at their computer
terminals lack the standards of appropriate behavior, the decision aids, the expert advice, the
market signals, and other forms of social support they would have if they met the same types of
problems regularly outside the laboratory. By depriving them of access to other people’s brains,
we make them appear clumsy, capricious, and less than fully rational. (Of course, many
individuals behave badly outside the laboratory—especially when they are on unfamiliar turf or
otherwise unsupported by a social network—and this qualifies as “irrational” by our definition if
they allow themselves to be easily exploited.)
Why are frequentist methods of statistics still so widely used, despite that fact that they are
demonstrably incoherent as a basis for decision making under uncertainty? In the last decade or
so, Bayesian methods have at last risen to prominence on the frontiers of statistics and artificial
intelligence as new computational tools such as Monte Carlo Markov Chain algorithms and
Bayesian networks have improved their tractability. But behind the front lines, introductory
textbooks and even scientific journal articles still speak of accepting or rejecting hypotheses on
the basis of their p-values. The persistence of such illogical methods can be understood by
observing that the results of statistical analysis are rarely plugged into well-formed decision
models where the effects of incoherence would be immediately felt. Rather, the analysis is
embedded in a social process that tends to compensate for its theoretical defects (or else obscure
them). The jargon of frequentist statistics serves as a lingua franca for communicating about
61
uncertainty in situations where the decision model is ambiguous: everyone misunderstands the
meaning of p-values, but at least they misunderstand them in the same way, and experienced
statisticians temper their use with rules of thumb that help to avoid embarrassing errors.
Bayesian methods place more of a focus—and more of a burden—on the individual who is doing
the modeling, and perhaps for that reason they have not yet fitted as well into the social process
of social-scientific inquiry. (Of course, the counterargument can also be made that a lot of bad
social science has managed to cloak itself in bad statistics.)
Why do individuals bother to vote, if their expected impact on the outcome of the election is less
than the trouble it takes? To frame the voting problem in this fashion is to overlook the entire
group decision process of which the casting of votes is only the culmination. The election
process not only aggregates but also serves to construct the preferences of the voters and to
manufacture candidates that will claim to satisfy those preferences, spurred on by interest groups
and news media. What happens on election day itself is often a mere formality, a ratification of
an outcome already determined by opinion polls. Candidate x or candidate y is “chosen” by the
final tally of votes, but what is more important is the campaign process through which the
political landscape has been shaped and the alternatives x and y have been conjured up.30 The
difference in expected utility that a given voter perceives between voting-for-x or voting-for-y or
not-voting-at-all on election day cannot be exactly quantified, nor is her voting strategy an
equilibrium of a game whose rules could have been written down in advance. Her own vote may
be a largely symbolic expression of her identity as a citizen. What matters is that by engaging in
a political process that unfolds gradually and somewhat unpredictably over time, the voters,
lobbyists, politicians, and reporters as a group cause action to be taken (for good or ill) on social
problems whose full dimensions they cannot comprehend by themselves. Of course this does
not absolve individuals of responsibility for trying to understand their economic and political
environment—because the aggregation of their incomplete mental models will eventually
determine the collective solution—but it does suggest the need for both humility and tolerance.
4.3 Implications for modeling
The theory of choice outlined here is broadly consistent with the normative ideals of optimization
and equilibrium that are central to standard rational choice theory, notwithstanding some
indeterminacies and unobservables. But it also reinforces many of the principal arguments of
rational choice critics, and its implications for modeling are distinctly different from those of the
standard theory in a number of important respects.
First, decision analysis should embrace the idea that decisions are inherently interactive rather
than celebrate the heroically rational individual. The aim of decision analysis should be to use
multiple brains to find creative solutions to complex problems, not merely satisfy the latent
preferences of a designated decision maker. Some of the brains may be multiple selves of the
decision maker, evoked by exercises in reframing, role-playing, and value-focused thinking. But
in most cases the individual’s colleagues, expert advisers, superiors, subordinates, role models,
This characterization of the election process is actually more reminiscent of “culturalist” or “interpretivist” views
than standard rational choice models in political science.
30
62
fellow consumers or investors—and decision analyst, if any—will also play a role as participants
in the decision, not mere catalysts or passive data sources. Waiting may sometimes help: it may
be advantageous to purchase an option to put off a decision until after some of its complexities
are resolved by the passage of time, and the value of the option may reside partly in the
opportunity to let unforeseen things happen. Of course, decision analysis practitioners already
know these things, but they consider them part of their art rather than their science. Indeed,
paradoxically, there is no role for a decision analyst in the standard rational choice theory—the
decision maker already implicitly knows what she ought to do. The view of choice developed
here suggests, to the contrary, that a second brain ought to be helpful, particularly one that has
broad experience with similar kinds of decisions or expert knowledge of the problem domain.
This may explain why software packages for decision-tree-analysis and expected-utility
maximization have not proved particularly helpful to individuals acting alone—even decision
theorists do not often use them in their private decisions.
Second, a novel approach to the modeling of games of strategy is suggested: let the common
knowledge assumptions be translated into gambles or trades that the players are willing to
publicly accept. Then impose a no-arbitrage condition on the resulting (observable) market,
rather than imposing an equilibrium concept on the players’ (unobservable) probabilities and
utilities. Of course, by Theorem 3, the two approaches are dual to each other, except that the
arbitrage-free equilibria will be subjective correlated equilibria rather than Nash equilibria and
the common prior distribution will be a risk-neutral distribution rather than anyone’s true
distribution. But more importantly, by operationalizing the common knowledge assumptions in
terms of material gambles or trades, the players are given the ability to tweak the rules of the
game. The question then becomes not only “how should this game be played?” but also “should
another game be played instead?” For example, two risk-averse players would never implement
a mixed strategy Nash equilibrium: if side bets were permitted, they would bet with each other
(or with an entrepreneurial observer) in such a way as to equalize their rankings of the outcomes,
thereby eliminating the need for randomization and perhaps even dissolving the game (Nau
1995c). The theory of games that emerges here is thus a theory of how to make the rules as well
as how to play once the rules have been fixed.
Concerning the study of economic and social institutions in which many individuals interact with
each other—households, firms, markets, and governments—the implication is even more
fundamental. To assume that it is “as if” every individual has a complete model of her
environment and is behaving optimally in it is to deny what is perhaps the most important reason
for them to interact, namely that their models are incomplete, and usually profoundly so. No
wonder this leads to paradoxes and puzzles. As Arrow (1987) observes: “If every agent has a
complete model of the economy, the hand running the economy is very visible indeed.” This is
not to suggest that mathematical models are unhelpful, but merely that plural analysis and highbandwidth data are likely to be needed to fully illuminate complex social or economic
phenomena. Painstaking field studies of institutions, rich data from real markets and polities, in
vivo social experimentation, lessons drawn from history and literature—all of which harness the
power of many people’s brains—may shed more light on human affairs than the exploration of a
system of equations with a small number of parameters. Certainly there is an important unifying
mathematical principle that underlies rational behavior in decisions, games, and markets: it turns
63
out to be the principle of no-arbitrage, once the knowledge assumptions are operationalized and
the dual side of the model is examined. But that principle has little predictive ability by itself:
the devil is still in the details.
Acknowledgements: I am grateful to Bob Clemen, Bart Lipman, Mark Machina, Peter Wakker,
and Andy Yates for helpful comments, but they are not responsible for the consequences. This
research is supported by the National Science Foundation under grant 98-09225 and by the Fuqua
School of Business.
References:
Allais, M. (1953) Le Comportement de l’Homme Rationnel devant le Risque: Critique des
Postulats et Axiomes de l’Ecole Américaine. Econometrica 21, 503—546.
Anscombe, F.J. & R.J. Aumann (1963) A Definition of Subjective Probability. Annals of
Mathematical Statistics 34, 199—205.
Arrow, K. J. (1951a) An Extension of the Basic Theorems of Classical Welfare Economics.
Proceedings of the Second Berkeley Symposium on Mathematical Statistics and
Probability. University of California Press
Arrow, K.J. (1951b) Social Choice and Individual Values. Wiley, New York.
Arrow, K.J. (1987) Economic Theory and the Hypothesis of Rationality. In J. Eatwell, M.
Milgrave, and P. Newman (eds.) The New Palgrave: A Dictionary of Economics,
MacMillan Press Limited, New York
Arrow, K.J. (1994) Methodological Individualism and Social Knowledge. American Economic
Review 84:2, 1—9
Arthur, W. B. (1994) Bounded Rationality and Inductive Reasoning (The El Farol Problem).
Amer. Econ. Rev. Papers and Proceedings 84
Aumann, R.J. (1974) Subjectivity and Correlation in Randomized Games. J. Math. Econ. 1
67—96.
Aumann, R.J. (1987) Correlated Equilibrium as an Expression of Bayesian Rationality.
Econometrica 55, 1—18.
Aumann, R.J. (1995) Backward Induction and Common Knowledge of Rationality. Games and
Economic Behavior 8:1, 6—19
64
Barkow, J., Cosmides, L. and J. Tooby (1992) The Adapted Mind : Evolutionary Psychology
and the Generation of Culture. Oxford University Press, New York
Bernheim, D.B. (1984) Rationalizable Strategic Behavior. Econometrica 52, 1007—1028.
Black, F. and M. Scholes (1973) The Pricing of Options and Corporate Liabilities. J. Political
Economy 81, 637—654
Binmore, K. (1994) Game theory and the Social Contract. Volume 1. Playing Fair. MIT Press,
Cambridge, Massachusetts
Brandenburger, A. and B. Nalebuff (1995) The Right Game: Use Game Theory to Shape
Strategy. Harvard Business Review 73:4, 57—71
Brodbeck, M., ed. (1968) Readings in the Philosophy of the Social Sciences. Macmillan, New
York
Camerer, C. and T.-H. Ho (1998) Experience-Weighted Attraction Learning In Normal Form
Games. Forthcoming, Econometrica
Chew, S.H. & L.G. Epstein (1989), A Unifying Approach to Axiomatic Non-Expected Utility
Theories. Journal of Economic Theory 49, 207—240.
Chew, S.H., L.G. Epstein, & P.P. Wakker (1993), A Unifying Approach to Axiomatic NonExpected Utility Theories: Correction and Comment. Journal of Economic Theory 59,
183—188.
Coleman, J.S. (1990) Foundations of Social Theory, Harvard University Press, Cambridge
Cook, K. and M. Levi, eds. (1990) The Limits of Rationality. University of Chicago Press
Cyert, R. and J. March (1963) A Behavioral Theory of the Firm. Prentice-Hall, Englewood
Cliffs, New Jersey
Dawkins, R. (1976) The selfish gene. Oxford University Press, New York
Dawkins, R. (1986) The Blind Watchmaker. Norton, New York
de Finetti, B. (1937) La Prévision: Ses Lois Logiques, Ses Sources Subjectives. Ann. Inst. Henri
Poincaré 7, 1—68. Translation reprinted in H.E. Kyburg and H.E. Smokler, eds. (1980)
Studies in Subjective Probability, 2nd ed., Robert Krieger, New York, 53—118
de Finetti, B. (1974) Theory of Probability, Vol. 1. Wiley, New York
Dennett, D. (1995) Darwin's Dangerous Idea : Evolution and the Meanings of Life. Simon &
Schuster, New York
Drèze, J. (1970) Market Allocation Under Uncertainty. European Econ. Rev. 2, 133—165
65
Drèze, J. (1987) Essays on Economic Decision under Uncertainty. Cambridge University Press,
London.
Edgeworth, F. Y. (1881) Mathematical Psychics: An Essay on the Application of Mathematics to
the Moral Sciences. Reprinted 1967, M. Kelley, NewYork.
Einhorn, H.J. & R.M. Hogarth (1981) Behavioral Decision Theory: Processes of Judgement and
Choice. Annual Review of Psychology 32, 53—88.
Ellsberg, D. (1961) Risk, Ambiguity and the Savage Axioms. Quarterly Journal of Economics
75, 643-669.
Elster, J., ed. (1986) Rational Choice. University Press, New York
Fishburn, P.C. (1988) Nonlinear Preference and Utility Theory. Johns Hopkins University Press,
Baltimore, MD
Ford, J., ed. (1992) Time, Expectations, and Uncertainty in Economics: Selected Essays of
G.L.S. Shackle. Edward Elgar Ltd., Hants, England
Friedman, J. (1995) The Rational Choice Controversy. Yale University Press, New Haven
Friedman, M. (1953) Essays in Positive Economics. University of Chicago Press
Garman, M. (1979) A Synthesis of the Pure Theory of Arbitrage. Working Paper No. 98.
Institute of Business and Economic Research, University of California, Berkeley
Genest, C. and J.V. Zidek (1986) Combining Probability Distributions: A Critique and
Annotated Bibliography. Statistical Science 1, 114—148
Gibbard, A. (1973) Manipulation of Voting Schemes: A General Result. Econometrica 41
587—601
Gigone, D. and R. Hastie (1997) Proper Analyis of the Accuracy of Group Judgments.
Psychological Bulletin 121:1 149—167
Giron, F.J., and S. Rios (1980) Quasi-Bayesian Behavior: A More Realistic Approach to
Decision Making? In J.M. Bernardo, M.H. DeGroot, D.V. Lindley, and A.F.M. Smith
(eds.), Bayesian Statistics. University Press, Valencia, Spain
Goldstein, M. (1983) The Prevision of a Prevision. J. Amer. Statist. Assoc. 78, 817—819
Goldstein, M. (1985) Temporal Coherence (with discussion). In J. Bernardo, M. DeGroot, D.
Lindley, and A. Smith (Eds.), Bayesian Statistics 2: Proceedings of the Second Valencia
International Meeting, (North-Holland, Amsterdam, 231—248.
Good, I. J. (1962) Subjective Probability as the Measure of a Non-Measurable Set. In E. Nagel,
P. Suppes, and A. Tarski (eds.) Logic, Methodology, and Philosophy of Science, Stanford
66
University Press. Reprinted in H.E. Kyburg and H.E. Smokler (eds.) (1980) Studies in
Subjective Probability, 2nd ed., Robert Krieger, New York, 133—146
Gode, D. and S. Sunder (1993) Allocative Efficiency of Markets with Zero-Intelligence Traders:
Market as a Partial Substitute for Individual Rationality. J. Political Economy 101:1,
119—37
Gould, S.J. (1989) Wonderful Life: The Burgess Shale and the Nature of History. W. W.
Norton, New York
Green, D.P. and I. Shapiro (1994) Pathologies of Rational Choice Theory. Yale University Press,
New Haven
Groves, T. and J. Ledyard (1987) Incentive Compatibility Since 1972. In T. Groves, R. Radner,
and S. Reiter (eds.), Information, incentives, and economic mechanisms: Essays in honor
of Leonid Hurwicz. University of Minnesota Press, Minneapolis. pp. 48—111
Hacking, I. (1967) Slightly More Realistic Personal Probability. Philosophy of Science 34,
311—325.
Hammond, P. (1997) Subjective Expected Utility and Non-cooperative Games. In S. Barberà, P.
Hammond and C. Seidl (eds.), Handbook of Utility Theory. Kluwer Academic
Publishers, Dordrecht
Harrison, J. and D. Kreps (1979) Martingales and Arbitrage in Multiperiod Securities Markets.
J. Econ. Theory 2:3, 381—408
Harsanyi, J.C. (1967,1968) Games With Incomplete Information Played By `Bayesian’ Players:
I-III. Management Science 14, 159—182, 320—334, 486—502.
Harsanyi, J.C. (1982) Subjective Probability and the Theory Of Games: Comments On Kadane
And Larkey’s Paper, And Rejoinder To Reply, Management Science 28, 120—124,
124—125.
Heap, S., M. Hollis, B. Lyons, R. Sugden, and A. Weale (1992) The Theory of Choice: A
Critical Guide. Blackwell, Oxford
Hill, G. (1982) Group Versus Individual Performance: Are N+1 Heads Better Than One?
Psychological Bulletin 91:3 517—539
Hogarth, R. and M. Reder, eds. (1987) Rational Choice: The Contrast Between Economics And
Psychology. University of Chicago Press
Homans, G. (1967) The Nature of Social Science. Harcourt, Brace, & World, New York
Howard, R.A. (1988) Decision Analysis: Practice and Promise. Management Science 34:6,
679—695
67
Howson. C., and P. Urbach (1989) Scientific Reasoning. Open Court, LaSalle, Illinois
Isenberg, D. (1984) How Senior Managers Think. Harvard Business Review, Nov-Dec, 80—90
Kadane, J.B. & R.L. Winkler (1988) Separating Probability Elicitation from Utilities. Journal of
the American Statistical Association 83, 357—363.
Kadane, J.B. and P.D. Larkey (1982), Subjective Probability and the Theory of Games, and
Reply, Management Science 28, 113—120, 124.
Kadane, J.B. and P.D. Larkey (1983), The Confusion of Is and Ought In Game Theoretic
Contexts, Management Science 29 1365—1379.
Kahneman, P. Slovic, & A. Tversky (1982, Eds.), Judgment under Uncertainty: Heuristics and
Biases, Cambridge University Press, Cambridge.
Keeney, R. and H. Raiffa (1976) Decisions with Multiple Objectives: Preferences and Value
Tradeoffs. Wiley, New York
Keeney, R. (1992) Value-Focused Thinking: A Path to Creative Decisionmaking. Harvard
University Press:, Cambridge, Massachusetts
Kerr, N., G. Kramer, and R. MacCoun (1996) Bias In Judgement: Comparing Individuals and
Groups. Psychological Review 103:4 687—719
Kirzner, I. (1992) The Meaning of Market Process: Essays in the Development of Modern
Austrian Economics. Routledge, London
Koopman, B.O. (1940) Axioms and Algebra of Intuitive Probability. Ann. Math. 41, 269—292
Kreps, D. (1988) Notes on the Theory of Choice. Westview Press, Boulder, Colorado
Lane, D, F. Malerba, R. Maxfield, and L. Orsenigo (1996) Choice and Action. J. Evolutionary
Economics 6:1, 43—76
Lintner, J. (1969) The Aggregation of Investors’ Diverse Judgments and Preferences in Purely
Competitive Security Markets. J. Financial & Quantitative Analysis 4:4, 346—400
Loewenstein, G. and J. Elster (1992) Choice Over Time. Russell Sage Foundation, New York
McClennen, E. F. (1990) Rationality and Dynamic Choice: Foundational Explorations.
Cambridge University Press, Cambridge.
Machina, M. (1982) Expected Utility Without the Independence Axiom. Econometrica 50,
227—323
Machina, M. (1983) Generalized Expected Utility Analysis and the Nature of Observed
Violations of the Independence Axiom. In B.P. Stigum and F. Wenstop, eds.,
68
Foundations of Utility Analysis and Risk Theory with Applications, Reidel, Dordrecht,
Holland, 263—293
Machina, M.J. (1987) Choice under Uncertainty: Problems Solved and Unsolved. Economic
Perspectives 1, 121—154.
March, J. (1994) A Primer on Decision Making: How Decisions Happen. Free Press, New York
March, J. and H. Simon (1993) Organizations, 2nd Ed. Blackwell, New York (1st ed. 1958, J.
Wiley, New York)
Merton, R. (1973) The Theory of Rational Option Pricing. Bell Journal of Economics and
Management Science 4, 141—183
Milgrom, P. and N. Stokey (1982) Information, Trade, and Common Knowledge. J. Econ.
Theory 26, 17—27
Modigliani, F. and M. Miller (1958) The Cost of Capital, Corporation Finance, and the Theory of
Investment. Amer. Econ. Rev. 48, 261—297
Moser, P., ed. (1990) Rationality in Action: Contemporary Approaches. Cambridge University
Press, New York
Munier, B. (1992) Market Uncertainty and Belief Formation with a Finite Number of Events. J.
Risk and Uncertainty 4:3, 233—250
Myerson, R. (1985) Bayesian Equilibrium and Incentive Compatibility. In: L. Hurwicz, D.
Schmeidler, and H. Sonnenschein, eds., Social Goals and Social Organization.
Cambridge University Press, Cambridge 229—259
Nau, R.F. (1989) Decision Analysis with Indeterminate or Incoherent Probabilities. Annals of
Operations Research 19 375—403
Nau, R.F. (1992a) Indeterminate Probabilities on Finite Sets. The Annals of Statistics 20:4
1737—1767
Nau, R.F. (1992b) Joint Coherence in Games of Incomplete Information. Management Science
38:3 374—387
Nau, R.F. (1995a) The Incoherence of Agreeing to Disagree.
239
Theory and Decision 39, 219—
Nau, R.F. (1995b) Coherent Decision Analysis with Inseparable Probabilities and Utilities.
Journal of Risk and Uncertainty 10 71-91
Nau, R.F. (1995c) Arbitrage-Free Correlated Equilibria. Working Paper, Fuqua School of
Business
69
Nau, R.F. (1999) A Generalization Of Pratt-Arrow Measure To Non-Expected-Utility
Preferences and Inseparable Probability And Utility. Working paper, Fuqua School of
Business and INSEAD
Nau, R.F. and K. F. McCardle (1990) Coherent Behavior in Noncooperative Games. Journal of
Economic Theory 50:2 424-444
Nau, R.F. and K.F. McCardle (1991) Arbitrage, Rationality, and Equilibrium. Theory and
Decision 31 199-240
Ordeshook, P. (1986) Game Theory and Political Theory: An Introduction. Cambridge
University Press, New York
Payne, J., J. Bettman and E. Johnson (1993) The Adaptive Decision Maker. Cambridge
University Press
Pearce, D.G. (1984) Rationalizable Strategic Behavior And The Problem Of Perfection.
Econometrica 54, 1029—1050.
Pinker, S. (1997) How the Mind Works. Norton, New York
Poundstone, W. (1992) Prisoner’s Dilemma. Doubleday, New York
Prelec, D. (1998) The Probability Weighting Function. Econometrica, forthcoming
Quiggin, J. (1982) A Theory of Anticipated Utility. Journal of Economic Behaviour and
Organization 3, 323—343
Rawls, J. (1971) A Theory of Justice. Harvard University Press, Cambridge MA
Rios Insua, D. (1990) Sensitivity Analysis in Multi-objective Decision Making. Springer-Verlag,
Berlin
Ross, S. (1976a) The Arbitrage Theory of Capital Asset Pricing. J. Econ. Theory 13, 341—360
Ross, S. (1976b) Risk, Return, and Arbitrage. In I. Friend and J. Bicksler, eds., Risk and Return
in Finance. Ballinger, Cambridge, Massachusetts
Ross, S. (1978) A Simple Approach to the Valuation of Risky Streams. J. Business 51:3, 453—
475
Ross, S. (1987) The Interrelations of Finance and Economics: Theoretical Perspectives. Amer.
Econ. Rev. 77:2, 29—34
Roth, A. (1995) Introduction to Experimental Economics. In J. Kagel and A. Roth, eds.,
Handbook of Experimental Economics. Princeton University Press
70
Roth, A. and I. Erev (1995) Learning in Extensive Form Games: Experimental Data and Simple
Dynamic Models in the Intermediate Term. Games and Economic Behavior 8 164-212
Rubinstein, M. (1976) The Valuation of Uncertain Income Streams and the Pricing of Options.
Bell J. Economics and Management Science 7, 407—425
Satterthwaite, M. (1975) Strategy-Proofness and Arrow’s Conditions: Existence and
Correspondence Theorems for Voting Procedures and Social Welfare Functions. J. Econ.
Theory 10 187—217
Savage, L.J. (1954) The Foundations of Statistics. Wiley, New York. (Second edition 1972,
Dover, New York)
Schervish, M.J., T. Seidenfeld, & J.B. Kadane (1990) State-Dependent Utilities. Journal of the
American Statistical Association 85, 840—847.
Schmeidler, D. (1989) Subjective Probability and Expected Utility without Additivity.
Econometrica 57, 571—587
Seidenfeld, T. (1988) Decision Theory Without “Independence” or Without “Ordering:” What’s
the Difference? Economics and Philosophy 4, 267—315
Seidenfeld, T., M.J. Schervish and J.B. Kadane (1998) A Representation of Partially Ordered
Preferences. The Annals of Statistics, forthcoming
Shafer, G. (1986) Savage Revisited (including comments), Statistical Science 1, 463-501
Simon, H. A. (1947) Administrative Behavior; A Study Of Decision-Making Processes In
Administrative Organizations. Macmillan Co., New York
Simon, H.A. (1955) A Behavioral Model of Rational Choice. Quart. J. Econ. 59, 99—118
Simon, H.A. (1976) From Substantive to Procedural Rationality. In S.J. Latsis, ed., Method and
Appraisal in Economics. Cambridge University Press, Cambridge. Reprinted in H. A.
Simon, Models of Bounded Rationality. MIT Press, Cambridge, Massachusetts
Simon, H.A. (1986) Rationality in Psychology and Economics. In Hogarth and Reder, pp. 25—
40
Simon, H.A. (1997) Models of Bounded Rationality, Vol. 3. MIT Press, Cambridge MA
Smith, C.A.B. (1961) Consistency in Statistical Inference and Decision. J. Roy. Stat. Soc. B 23,
1—25
Smith, J.E. and R.F. Nau (1995) Valuing Risky Projects: Option Pricing Theory and Decision
Analysis. Management Science 41:5, 795-816
71
Sugden, R. (1991) Rational Choice: A Survey of Contributions from Economics and Philosophy.
Economic Journal 101, 751-785
Suppes, P. (1974) The Measurement of Belief. J. Roy. Stat. Soc. B 36, 160—175
Tversky, A. & C. Fox (1995) Weighing Risk and Uncertainty. Psychological Review 102 269—
283
Varian, H. (1987) The Arbitrage Principle in Financial Economics. J. Economic Perspectives 1,
55—72.
von Neumann, J. & O. Morgenstern (1947) Theory of Games and Economic Behavior. Princeton
University Press, Princeton NJ
von Winterfeldt, D. and W. Edwards (1986) Decision Analysis and Behavioral Research.
Cambridge University Press
Vromen, Jack J. (1995) Economic Evolution : An Enquiry into the Foundations of New
Institutional Economics. Routledge, London/New York
Walley, P. (1991) Statistical Reasoning with Imprecise Probabilities. Chapman and Hall,
London
Wakker, P.P. (1989) Additive Representations of Preferences: A New Foundation of Decision
Analysis. Kluwer Academic Publishers, Dordrecht.
Watkins, J.W.N. (1957) Historical Explanation in the Social Sciences. British J. for the
Philosophy of Science 8 104—117. Reprinted in Brodbeck (1968)
Wu, G. & R. Gonzalez (1996) Curvature of the Probability Weighting Function. Management
Science 42 1676-1690.
Yaari, M.E. (1987) The Dual Theory of Choice under Risk. Econometrica 55, 95—115.
72
Download